Stephen Hawking The Selling of A Scientific Celebrity - Charles Seife
Stephen Hawking The Selling of A Scientific Celebrity - Charles Seife
Stephen Hawking The Selling of A Scientific Celebrity - Charles Seife
Hachette Book Group supports the right to free expression and the value of
copyright. The purpose of copyright is to encourage writers and artists to
produce the creative works that enrich our culture.
Basic Books
Hachette Book Group
1290 Avenue of the Americas, New York, NY 10104
www.basicbooks.com
The Hachette Speakers Bureau provides a wide range of authors for speaking
events. To find out more, go to www.hachettespeakersbureau.com or call
(866) 376-6591.
The publisher is not responsible for websites (or their content) that are not
owned by the publisher.
E3-20210311-JV-NF-ORI
CONTENTS
Cover
Title Page
Copyright
Prologue
PART I RINGDOWN
PART II IMPACT
Acknowledgments
Discover More
About the Author
Also By Charles Seife
Notes
Explore book giveaways, sneak peeks, deals, and more.
The Daily Mail loved Stephen Hawking far more than Stephen Hawking
loved the Daily Mail. Even by UK tabloid standards, the Mail’s science
coverage was either laughable or infuriating, depending on your point of
view. The paper’s pages were always chock-full of headlines about
scientific research—research often hyped by the Mail’s writers almost
beyond the point of recognition. All the better to grab the attention of an
audience.
Nobody could grab an audience like Stephen Hawking, so his name
regularly graced the tabloid pages. Usually not in a flattering light. The
professor was typically either a harbinger of doom—warning of imminent
death due to global warming, robot rebellion, alien invasion, or other
catastrophes—or he was at the center of some sort of scandal about his sex
life or his marriages or abuse allegations. But in early 2018, just before
Hawking’s death, the Daily Mail broke new ground.
“Has Stephen Hawking Been Replaced with a ‘Puppet’?” the headline
read. “Conspiracy Theorists Claim the REAL Professor Is DEAD and a
‘Puppet’ Has Taken His Place—and Reveal the SIX Clues That Support the
Idea.”1
In a surprisingly long and detailed article, the tabloid set out evidence that
sometime in the mid-1980s, the esteemed physicist had been replaced with
an impostor. As outrageous as this theory sounds, the Mail explained how
supposed anomalies in his appearance as he aged (particularly the look of his
teeth), his unexpectedly long survival with a disease that typically kills in a
couple of years, and a number of other clues suggested that the original
Stephen Hawking had died and that a facsimile had been foisted on the
public. “The voice we hear,” said the article, “is the result of NASA
astrophysicists typing information into a computer—information they want…
to push on a gullible and unsuspecting public, fans of Hawking who hang on
to—what they believe to be his—every word.”
Even for the bizarre parallel reality conjured by tabloid writers, this was
way out there. They had ventured into this sort of territory only once before,
almost exactly fifty years prior. In 1969, the tabloid circuit lit up with rumors
that the Beatles’ Paul McCartney had been killed in an auto accident and
replaced with a doppelgänger.
However, comparing Stephen Hawking to Paul McCartney doesn’t quite
capture the nature of Hawking’s singular celebrity. Throughout all of history,
there might be perhaps three or four scientists whose fame and renown
among the public could be compared to Hawking’s: Einstein, Newton,
Galileo—maybe Darwin. For the media and for the public at large, Hawking
had become the ultimate symbol of the triumph of the mind. He was the
world’s smartest man, an unmatched brain who spent his time unraveling the
deepest mysteries of the universe.
The Mail’s suggestion that Hawking had been replaced by a simulacrum
was just the most extreme and absurd version of how the press and the public
had portrayed Hawking for decades. The professor’s image had been built
into a towering contradiction: On one hand, Hawking appeared to the world
to be something more than human, his mind so transcendent that he was in a
class by himself. He inhabited an intellectual plane above the realm of
normal humanity. Yet, on the other hand, he could be treated as a nearly
inanimate object. Hawking suffered from a neurological disease that slowly
robbed him of the ability to move of his own volition, to speak except
through a computer-generated voice. It was a short leap for a thoughtless
person to imagine Hawking to be artificial, to be some sort of technology-
assisted homunculus rather than a real human. As the Daily Mail so rudely
put it, it was not even always possible to tell whether the voice emanating
from his computer was truly under the control of the being sitting in the
wheelchair.i
By the time he died in 2018, it was almost impossible to discern Hawking
the human underneath the layers of accumulated symbolism; from the public’s
point of view, he had become a caricature rather than a living person. Even
though everyone who knew Hawking has described him as one of the most
stubborn, willful people they’ve ever met, it was incredibly hard to
distinguish his true will, to sense the authentic being underneath the public
persona.
To understand Stephen Hawking, one has to turn back the clock. During
the last third of his life, Hawking was firmly entrenched as the world’s most
celebrated living scientist, yet his actual scientific contributions were more
or less irrelevant to his fame. Though he was a regular fixture in the media,
the press attention wasn’t usually related to his science. Hawking’s research
during the years of his greatest popularity would be largely discounted and
have little lasting impact on the world of physics. He was like a collapsed
star: space around him glowed brightly with his energy, but at core he was
but a faint reflection of what he once had been.
Not long before, Hawking had been a supernova. The middle third of his
life was a spectacular and brilliant transformation. Over the course of two
decades, he transmogrified himself from a fairly obscure physicist, laboring
with his colleagues (and his rivals) to understand the conditions at the very
beginning of the universe, into an international celebrity. Into the world’s
smartest man. Into the scientific equivalent of The Beatles. It was a
metamorphosis at once immensely satisfying and intensely painful. By the
time it was complete, Hawking had broken with much of his past and
constructed a myth to replace it.
Only in the first third of Hawking’s life—before he achieved his status
and fame—does the real human being behind the legend begin to appear. The
backward flow of narrative time slowly restores Hawking to his primeval
brilliance. By traveling back to Hawking’s youth, one can understand how he
came to the key scientific insights upon which he built his name. One can
discern the roots of his need to become a famous communicator of science.
And one can understand the mortal fears of a young man racing to establish a
legacy—and a family—as a deadly disease appeared poised to claim him at
any moment.
Unlike a scientific discovery, which becomes easier to understand as time
moves forward and as more and more researchers contribute their
knowledge, the life of Stephen Hawking becomes clearer as time moves
backward, as the accumulated layers of celebrity and legend are stripped
away. In the end, Stephen Hawking the human being becomes something very
different from the Stephen Hawking so beloved by the public.
Public Hawking was famous for being the world’s smartest man, the apex
of scientific intellectualism. Human Hawking was brilliant, but he knew he
was surrounded by equally brilliant peers who labored in semi-obscurity.
Public Hawking was the world’s greatest science communicator. Hawking
the human had more difficulty communicating than just about any other person
on the planet; by the time he was famous, he could only express a few words
per minute, if that. Public Hawking stoically shrugged off his physical
ailments as a mere inconvenience. Human Hawking’s disability had, quite
naturally, shaped every part of his existence: his outlook, his science, his
family life, and eventually his fame. To the public, everything Hawking did
was extraordinary and different and courageous—it was a spectacle when he
spoke, when he ate, when he danced, when he worked, when he loved. To
Hawking, there was no courage in merely being himself.
Even Hawking’s colleagues and rivals had to take pains to distinguish the
human from the legend. “I’m not speaking of him as a pure intellectual who
rides through the universe on his magic wheelchair,” says Leonard Susskind,
a physicist at Stanford University who battled Hawking about the properties
of black holes. “I’m speaking of him as a human being. You know, none of us
were ever really able to know him.”2
Turn back the clock, and what emerges is a real human: petulant, arrogant,
and callous as well as warm, witty, and brilliant. Complex. Fascinating.
Singular.
What emerges is Stephen Hawking.
Footnote
i The public treasured any encounter with Hawking that seemed to express
his real, unfiltered personality. For example, he was notorious for his
bullheaded indifference to other vehicles when driving his wheelchair on the
streets around Cambridge University—and people found this utterly
charming. When Hawking died in 2018, one of the viral discussions on
Twitter was a long thread filled with stories of people who nearly ran him
over with their cars.
PART I
RINGDOWN
Only thrice in the past hundred years has a scientist been buried in
Westminster Abbey. There was Ernest Rutherford, who figured out the
structure of the atom; there was J. J. Thomson, who discovered the electron;
and then there was Stephen Hawking.
On June 15, 2018, Hawking’s ashes were interred in the floor of the
cathedral, laid to rest underneath a slate-black stone just a few feet away
from the graves of Isaac Newton and Charles Darwin.
Hawking publicly disavowed any comparison to Newton—whenever
anyone made such a suggestion, he would rubbish it as “media hype.” Yet the
public loved to make the connection: like Newton, Hawking was the most
famous physicist of his day; at Cambridge, Hawking occupied the Lucasian
chair, the very office that Newton had held three hundred years before him;
Hawking and Newton had both devoted much of their lives to understanding
the mysteries of the gravitational force. Even in death, Hawking could not
escape the association with Newton. Not only are the two scientists buried
within a few strides of one another, but they have the same epitaph. Newton’s
black tombstone is featureless but for an inscription in Latin: “Hic depositum
est quod mortale fuit Isaaci Newtoni.” Hawking’s bears the same words in
English, with his name substituted for Newton’s: “Here lies what was mortal
of Stephen Hawking.”1
Though smaller than Newton’s gravestone, Hawking’s is more elaborate.
The epitaph curves gently around a set of swirls engraved into the slate,
swirls that seem to move toward an elliptical void: clouds of gas falling into
the maw of a black hole. To the left, there is an equation whose letters
seemingly defy gravity:
T = ħc3/8πGMk
Almost nobody who visits the gravestone understands what those symbols
mean. But to Stephen Hawking, that equation was the key to transcending the
mortal.
Until he died in 2018, Hawking was one of the most recognized human
beings on the planet—and probably the easiest to spot. Almost unable to
move his body, seated awkwardly in his motorized wheelchair and
accompanied by an entourage of nurses, he was unable to go anywhere
incognito. Not that he cared to.
The public adored Hawking without knowing precisely why. Einstein had
his theory of relativity and Newton his universal gravitation, but the vast
majority of people who admired Hawking knew little about what he had done
to deserve his reputation. Nor did they understand why, in the press, he was
always compared to Einstein or Newton, a comparison he modestly rejected
but at the same time worked very hard to cultivate. And even those who had a
glimmer of Stephen Hawking’s science saw only a tiny fraction of what made
Hawking Hawking. For he was not just Hawking the physicist, Hawking the
celebrity; he was Hawking the showman, Hawking the husband and father,
Hawking the symbol.
These facets warred among themselves: the very moment Hawking
achieved celebrity, his marriage collapsed and his family shattered. Hawking
the human depended on his students to be caretakers and nursemaids even as
Hawking the physicist wanted to groom them to be his intellectual offspring.
He was perhaps the most celebrated communicator of physics in the world,
yet he had extraordinary difficulty making himself understood. Even the most
straightforward-seeming element of his persona, his ability as a first-class
physicist, is much more complicated than it seems at first. Scientists viewed
Hawking as a mind of a very high order—but at the same time, many rolled
their eyes at some of his later work, trashing it as all but worthless. The real
Hawking lies underneath this complex skein of tangled and contradictory
narratives.
As with the black holes that he studied, there are incredible forces that
prevent outsiders from glimpsing Hawking’s inner self. But there is a real
person who lies beyond the event horizon of his celebrity.
That singularity contained legion: an important scientist whose
importance is almost universally misunderstood; a person who suffered
deeply and also caused deep suffering; a celebrity scientist who broke the
mold of his forebears and fundamentally changed the concept of a scientific
celebrity.
Most people who know anything of Hawking are blinded by a flash from
his life, an image of a tumultuous decade—from 1980 through 1990—when
he transformed himself from a well-respected but obscure scientist in a
neglected corner of physics into one of the most recognized names on the
planet. But like a supernova that briefly outshines the rest of its host galaxy,
Hawking’s celebrity both attracted attention and distracted it—
simultaneously inviting the gaze of countless millions and hiding the star
itself, a quivering and naked object that shed everything that once clung to it.
Unerring he was not, but Stephen Hawking had made it his profession to
try to understand the beginning and the end of the universe. When he began
his research in the early 1960s, his field, cosmology, was a sleepy
backwater, an area of study that hadn’t seen a substantial advance for
decades. By the time he died, it was arguably the most exciting field in
physics, an area that was (and still is) generating Nobel Prize after Nobel
Prize for transforming our understanding of how the universe came to be.
Hawking’s first substantial piece of research was an important discovery
about the beginning of the universe. At the time, in 1965, there were two
competing models of how the cosmos was born: whether it was eternally
renewing itself, or whether it had been born in a gigantic explosion now
known as the Big Bang. For his PhD thesis, Hawking proved that if the
universe began with a Big Bang, then it had to have started as a singularity: a
point where the laws of physics no longer make sense, an infinitesimal but
infinite blemish on the fabric of space and time. A place where mathematics
itself breaks down. This was a stunning insight; if one believed in the Big
Bang, one had to accept that the laws of physics as we know them are
inadequate to describe the birth of our cosmos. This idea—now known as the
singularity theorem—ignited Hawking’s career.
As he grew in confidence and stature, Hawking became a key figure in
solidifying what is now the dominant theory of how the very early universe
expanded, a theory known as inflation. But Hawking himself thought his most
important contribution to cosmology was his work on an ambitious, radical,
and controversial theory in which he attempted to calculate the quantum-
mechanical “wavefunction of the universe.” Not only did Hawking believe
that his theory described the very starting point of space and time in our
cosmos, but he was convinced that it did so in a way that made God
unnecessary. “What place, then, for a creator?” he asked, much to the chagrin
of many theologians (and some scientists) around the world.2
However, Hawking’s most important scientific work wasn’t about the
birth of our universe or its wavefunction, but about a different kind of
singularity: the singularity at the heart of a black hole. Hawking devoted
much of his life to understanding how these mysterious objects behaved, and
in the most important moment in his scientific lifetime, he realized they had a
bizarre property that nobody else had imagined they might possess.
Black holes are astronomical objects whose gravitational attraction is so
powerful that nothing venturing too close can escape—not even light. Black
holes are born when a large star dies; when the fusion engine at its heart
shuts down, it collapses under the force of its own gravity. In a fraction of a
second, the entire weight of a star bears down on itself, first crushing the
matter into an undifferentiated glob of atoms, then crushing the atoms
themselves—and then, finally, it becomes a singularity. Because the
gravitational pull around the collapsed star is so great, nothing can venture
close to that singularity and escape to tell the tale. It’s as if the former star is
now surrounded by an invisible shroud marking the point of no return: cross
this so-called event horizon, and you are doomed, unable to return home,
destined to fall into the black hole no matter how hard you struggle.
Because black holes swallow light, they are as black as black can be;
they are the ultimate absorbers, gobbling any illumination rather than
reflecting it. But in the 1970s, Hawking had a surprising realization: black
holes aren’t really perfectly black after all. They radiate particles, including
light particles, in all directions. Under most circumstances, this radiation—
now known as Hawking radiation—is incredibly weak, far too weak to be
detected at any reasonable distance. However, just the fact that the radiation
exists had some profound implications. Because if a black hole radiates
energy, this means it must eventually evaporate—explode—in a burst of
radiation. This, in turn, implies that the matter and energy swallowed by the
black hole must eventually be released. And, as Hawking was the first to
realize, the release of that matter and energy leads to a seemingly
irreconcilable clash between the two mainstays of modern physics:
Einstein’s theory of relativity and quantum theory. The discovery of Hawking
radiation not only upended the conventional wisdom about black holes, but
seemed like a major milestone in the quest to resolve the conflict between the
two theories. Perhaps he could even replace them with an overarching
“theory of everything.”
“I would say, in retrospect, [Hawking] has three great contributions to
science. One is the singularity theorems,” says John Preskill, a physicist and
friend of Hawking’s. “And one is the idea about the wavefunction of the
universe. But the most important, by far, is the discovery of Hawking
radiation and its implications.”3
The equation inscribed on Hawking’s tombstone is the main formula for
Hawking radiation—the temperature of a black hole as a function of its mass,
which, in turn, dictates the amount and type of radiation that it emits—all
superimposed on the black hole that it describes.
As the grand organ in Westminster Abbey swelled, hundreds of voices sang
in unison to an old English melody:
Almost automatically, visitors would imbue Hawking not just with deep
wisdom, but also with childlike simplicity. He was becoming a guru, a
symbol. A metaphor—and one that was almost too perfect. Immobile in his
wheelchair, Hawking was a being of pure intellect, a man whose powerful
mind allowed him to travel to domains where nothing else in the universe
could venture.
Hawking was well aware that the myth, the archetype, was powerful
enough to swallow all traces of the real human being underneath. “That
Stephen is some sort of pure mind because of his disability, I think that hurt
him a lot,” says Christophe Galfard, one of the many PhD students Hawking
advised over the years. “The man as a scientist, not just the image of the
scientist, was somehow diluted.” So Hawking struggled to prevent that from
happening. In his home life, Hawking refused to discuss, much less make
concessions to, his disability, almost to a pathological extent; this obstinacy
became a sore spot in his first marriage. As a physicist, he tried to produce
ideas of such depth and importance that his physical impediments would be
seen as irrelevant. “I would like to be thought of as a scientist who just
happens to be disabled, rather than as a disabled scientist,” he would say.
Yet throughout his life, Hawking suspected that people saw his disability as a
mitigating factor, a reason to judge him differently from other physicists. Or
worse, as something that would come to define him. His fears were well
founded.6
The disability was central to Hawking the Metaphor, even if it was in
many ways peripheral to Hawking the Human. And as much as he wanted
people to see beyond his condition, Hawking realized, to his chagrin, that his
disability was right at the core of his public persona.
RIPPLES (2014–2017)
W hen the alarm clock went off at 2:40 a.m., Barry Barish swallowed his
disappointment. “I assumed they had passed us over,” he later recalled. But
then his cellphone rang.1
The early-morning phone call is one of the clichés about getting a Nobel
Prize. A scientist is supposed to awaken, blearily, to the realization that he or
she is suddenly going to be an instant celebrity. It’s supposed to be stunning,
humbling—and above all, it’s supposed to be a surprise. The first Monday
evening in October is a restless night for many physicists of great renown,
but never for a moment does even the greatest-egoed among them go to bed
fully expecting to win the Nobel the next morning.
Except this time. October 3, 2017.
Two years earlier, a novel telescope had made a discovery so important
that a Nobel was not only assured, but would come at the earliest possible
opportunity. The only question was how the prize would be divvied up.
Though there were hundreds upon hundreds of people who had worked on
the telescope and its observatory, the rule was that a prize could be split at
most three ways.
Barish, the director of the observatory, went to bed confident. Rainer
Weiss at the Massachusetts Institute of Technology (MIT), who had spent
decades designing the machine, was a bit more humble; he went to bed that
evening thinking he had only a 20 percent chance of waking up to a Nobel.
The third person who fell asleep that night in anticipation of a Nobel the next
morning was California Institute of Technology (Caltech) theorist Kip
Thorne, Hawking’s colleague and close friend for five decades.2
Like Hawking, Thorne had devoted his life to studying black holes,
gravity, and time, and this new telescope was about to shed new light on
precisely these subjects. For the telescope that was about to win Thorne,
Barish, and Weiss the Nobel Prize was not the ordinary sort of telescope that
gathers light from distant stars. Instead, this telescope, the Laser
Interferometer Gravitational-wave Observatory (LIGO), was a tool designed
to detect not light, but gravitational waves from colliding black holes. And
with this tool, Thorne was able to begin testing the theories that Hawking,
Thorne, and other physicists had developed in the late 1960s and early 1970s
—a time so rich in theoretical discovery that Thorne dubbed it the Golden
Age of Black Holes. “Among the nicest features of the Golden Age was the
way we all built on each other’s work,” Thorne wrote. “Hawking laid the
foundations, and one after another his compatriots built an edifice upon
them.”3
With his Nobel, Thorne belatedly fulfilled a promise that he had made on
Hawking’s sixtieth birthday fifteen years earlier: “I’m afraid it is more in the
form of a promissory note than a concrete physics result,” he had said. “Your
birthday gift is that our gravitational-wave detectors [including LIGO] will
test your Golden-Age black-hole predictions, and they will begin to do so
well before your 70th birthday. Happy Birthday, Stephen!”4
By the time he reached his seventies, Stephen Hawking was the world’s
most famous living scientist, and had been for several decades. His first
guest appearance on the animated show The Simpsons—a reliable indicator
of apex pop-culture status—had happened nearly twenty years prior. And
Hawking’s most important research was two generations in the past.
In the last decades of his life, Stephen Hawking’s fame was not really that
of a scientist, but of a cultural icon. Though his celebrity was mantled in
science, science had become almost incidental to Hawking’s notoriety. It was
mostly irrelevant whether his latest pronouncements about physics were
valuable or not; just the fact that he would make them every so often was
more than enough to maintain his status as an icon. The public simply didn’t
care all that much about the details of Hawking’s scientific achievements or
ideas. Yet Hawking wanted to be famous for his physics, not for his
personality, or his condition, or anything else.
In the late 2010s, Hawking had hope that this would finally change.
Hawking’s work of thirty and forty years prior had suddenly become hot
again. The most exciting physics of the day had to do with gravitational
waves and black holes, areas of knowledge where Hawking’s scientific
work had had the most profound effect. After years upon years of labor,
physicists around the world (including Hawking’s best friend, Kip Thorne)
were finally getting results from experiments that held the promise of testing
a number of Hawking’s decades-old predictions. If they confirmed some of
his theoretical work, Hawking might finally achieve his wish to be known as
a brilliant scientist first and a celebrity second. However, that dream would
come to naught if he were left on the periphery when the Nobel committee
came calling.
Gμν = (8πG/c4)Tμν
Just shy of one hundred years after Einstein’s rise to become the public face
of science, Stephen Hawking was reigning in his place as the premier
scientific celebrity. But the expectations of celebrity had changed a little bit
in the interim.
“Analyzing data since the ’66 World Cup, I have answered two of the
biggest questions tormenting fans,” Hawking told a gathered crowd of
journalists in the basement of London’s Savoy Hotel in 2014. “One, what are
the optimal conditions for England’s success, and two, how do you score in a
penalty shootout?”6
The event was sponsored by Paddy Power, a Dublin-based bookmaker,
best known for chasing publicity by offering bizarre attention-getting bets.
(“As the oil disaster in the Gulf of Mexico enters its second month with little
or no sign of abating,” the company announced shortly after the Deepwater
Horizon accident in 2010, “leading betting outfit Paddy Power are taking
bets on the first species to become extinct as a direct result of the spill. Top
of the bookies list at odds of 4/5 is the already critically endangered Kemp’s
Ridley Turtle.”7) This time, Paddy Power was trying a different way of
getting some attention.
“The technique I have used is called general logistic regression
modeling,” Hawking announced. But Hawking’s “analysis” was anything but
scientific. “Our chance of triumph can be worked out by looking at a number
of variables. Statistically, England’s red kit [uniform] is more successful.”
For some reason, the British press seems inordinately fond of marketing
ploys dressed up as nonsensical mathematical formulae—a formula for the
perfect pizza to help drive sales for a pizza chain, an equation for the most
miserable day of the year to encourage Britons to purchase a weekend
getaway package from a certain travel agent, the formula for the perfect
pancake to sell a supermarket chain’s nonstick frying pans, and the like. To
give the formula some credibility, the sponsor typically shops around for a
scientist or a mathematician who’s willing to accept a fistful of cash in return
for lending his or her name to the whole silly endeavor. Typically, these
scientists don’t have enough of a reputation to damage it by producing
nonsensical equations. “All are commissioned by companies as PR stunts
and their value ends there,” a science journalist wrote in The Guardian.
“They are overwhelmingly drawn up by scientists whose names are unknown
to any Nobel committee.”8
This time was clearly different.
In his pitch to the press, Hawking seemed to be kidding around. (“As we
say in science,” he intoned, “England couldn’t hit a cow’s arse with a
banjo.”) Nevertheless, it was surprising that such an eminent scientist would
lend out his name for such a ridiculous publicity stunt. It was a shock even to
Paddy Power. A Paddy Power spokesperson later admitted that he never
expected Hawking to agree to the bookmaker’s request. “We thought there
was a one percent chance he’d say yes,” the spokesperson said. “But he did.
I was totally surprised.”9
When journalists asked Paddy Power representatives how much they had
paid Hawking for his services, they got no answer. However, Hawking
reportedly “said he split the fee between two charities, one devoted to saving
children in Syria, and the other to motor neurone disease, the condition
Hawking was diagnosed with as a student.”10
It’s not like the media ignored Hawking. Far from it. It’s just that the
headlines were almost never about his science. They were about his
pronouncements—or his personal life. No other scientist, not even Einstein,
had a life story that so captivated the public, or had so many films made
about it.
“Focus Features’ much-awaited ‘The Theory of Everything’ world
premiered at Toronto on Sunday night to the most rapturous standing ovation
of the festival so far,” gushed the Hollywood trade magazine Variety. “The
Stephen Hawking biopic starring Eddie Redmayne and Felicity Jones (as his
wife Jane Hawking) left not a dry eye in the house.”18
It was not the first movie about Hawking, but it was arguably the best
received. (It soon won the gangly, freckle-faced Redmayne an Oscar for his
portrayal of the physicist as he grappled with the disease that robbed him of
his ability to move.) At its heart, though, The Theory of Everything was a
tear-jerker of a love story, one that Hawking wryly described as “broadly
true.”19
The story begins with the young Stephen as a geeky prodigy—a healthy
one—beginning his studies at Cambridge University. Soon after meeting the
love of his life, Jane Wilde, he is diagnosed with motor neurone disease and
given two years to live. Naturally, he sinks into an angry depression. Jane’s
love pulls him back from the slough of despond, and Stephen decides to make
the best of the time he has left by studying time itself.
The love story is star-crossed from the start. Jane and Stephen spar
throughout the movie about religion; a devout Anglican, she interprets his
work—his attempts to “prove with a single equation that time had a
beginning… one simple, elegant equation that will explain everything”—as
alternately affirming and denying the existence of the creator who is so dear
to her. And as Stephen’s disease gets progressively worse, Jane turns to a
widowed choirmaster, Jonathan Hellyer Jones, to help maintain the
household and to act as a surrogate father for Stephen’s children. Jane begins
to fall for Jones, and when Stephen tells his wife that he “won’t object” to
Jones’ presence in the household, Stephen’s tortured expression makes clear
the true nature of what that permission truly meant.20
But Jane refuses to act upon her feelings; the movie implies that her
relationship with Jones is entirely chaste despite their mutual attraction. Only
after Stephen falls for his nurse—leading to a weepy and poetic mutual
parting of the ways—can Jane finally act on her suppressed feelings. Yet the
two remain friends, and the movie ends with the pair looking lovingly at each
other, hand in hand, as their three happy children play nearby.
Except for the few moments of despair and pain, Redmayne’s Hawking
always sports a twisted, wicked grin. He is affable and sympathetic even as
he leaves his wife of twenty-four years for the nurse who seduced him.
Felicity Jones’ Jane seems peevish; the actress’ attempts to look stoic and
determined often come across as being annoyed at her husband’s disability.
This was almost certainly not what Jane had in mind when she agreed to sell
the movie rights for her book.
The Theory of Everything was based on Jane Hawking’s five-hundred-
page memoir, Travelling to Infinity: My Life with Stephen Hawking, which
in turn was a reworked version of her earlier six-hundred-page tell-all,
Music to Move the Stars. In her books, Jane was able to control the
narrative. She seemed somewhat shocked when the movie version didn’t tell
the story quite how she expected.
“The film really only shows part of our lives in Cambridge,” Jane
Hawking told The Guardian, explaining that the movie didn’t give a sense of
how much strain it was to take care of Stephen, especially with all the
traveling he had to do. (She was rebuffed when she asked for the insertion of
a video montage of frenzied packing and stowing and preparing for a trip.)
Nor did she appreciate that she “didn’t seem to have any friends or relations
at all” in the movie. “I knew that if there were mistakes in the film that they
were going to be immortalised, which they have been,” she said, adding, “I
found that very irritating and I didn’t want it to happen. Don’t ever believe
what you see in films.”21
Though the movie was based on his ex-wife’s book, rather than his own
autobiography, Hawking helped a great deal with the production. He spent
time with Redmayne, and even lent the production crew the use of his
distinctive robotic voice. “We’d been using this synthetic version of his
voice which this company had drawn up for us,” Redmayne told Empire, a
movie magazine. “At the end [of the early screening] he gave us the copyright
to use his actual voice.”22 vi
Jane, as the author of the book that inspired the film, naturally got pride of
place in the credits at the end of the movie; she is mentioned right after the
director, producers, and screenplay writer, and in an equally large font.23vii
Then come credits for the actors and crew and thank-yous and
acknowledgments for dozens more, including various artists, the locations
where the filming took place, and the organizations that gave permission to
reprint images in the movie. Even Jane’s parents, who had been dead for
years before the filming began, got a special thank you.
Nowhere acknowledged in the credits at all: Stephen Hawking.
Stephen had his own film projects, though nothing even close to the size of
what his ex-wife had been working on. (The Theory of Everything grossed
more than $120 million worldwide.24 viii) The professor and his robotic
voice were a regular staple on cable-TV science shows.
Stem Cell Universe with Stephen Hawking, which aired on the
Discovery Science Channel in April 2014, opened with a shot of the
physicist in his wheelchair, his image superimposed upon a luminous spiral
galaxy that wheels slowly around him as he narrates. “I have spent my life
exploring the mysteries of the cosmos, but there’s another universe that
fascinates me. The one hidden inside our bodies.” With that, the galaxy
suddenly contracts, collapsing into Hawking’s midriff. And then, with a bang,
a disk of little glowing indistinct blobs begins orbiting the physicist anew.
“Our own personal galaxies of cells. Today we are on the brink of a new age
in medicine. An age where we will be able to heal our bodies of any illness,
all because of cells inside us which have special powers.”25
Hawking can be forgiven for his hyperbole; after all, he is a cosmologist,
not a biologist or a physician. The clumsy galaxies-to-cells sequence
attempts to paper over Hawking’s lack of expertise—and perhaps even real
familiarity with—stem cell research. It doesn’t matter; Hawking is science
incarnate. His mere presence signals to the audience that what follows is
serious, cutting-edge research. Hawking’s six-part series, Science of the
Future, which aired on the National Geographic Channel in 2014, dealt with
such subjects as virtual reality, robots, urban design, and military technology
—nothing anywhere close to the physicist’s areas of study. Hawking had
granted TV producers his voice and his name to give the show credibility;
there wasn’t much else he could (or needed to) contribute.
Unique among celebrities of the day, Hawking could, quite literally, lend
his voice to a production. In the mid-1980s, when doctors performed a
tracheostomy to save his life, Hawking lost the use of his larynx and his
ability to speak. A team of engineers and software designers rigged his
wheelchair with a computer system that he could operate despite his ever-
diminishing muscle control. Embedded within that system was a speech
synthesizer. Hawking could slowly compose a sentence, and then, with one
final twitch of his muscles, send the text to the speech box, which would then
attempt to pronounce the words that Hawking had keyed in.
Hawking’s voice was truly disembodied; it resided in a little computer
that could—and did—act independently of its master.
Hawking sometimes let others compose sentences for him, which could
then be uploaded into his wheelchair computer. Hawking could then edit
them—or not—as he saw fit. It was far more efficient than having to
laboriously construct his own sentences from scratch on his computer.ix The
act of a twitch, a muscular assent, sent those foreign words through his own
extended body, to be pronounced through his own voicebox, and become his
own.
The human in the wheelchair didn’t even have to be present. By the time
Hawking lent out his voice to the producers of The Theory of Everything, he
had, on occasion, been allowing filmmakers to use his speech synthesizer for
nearly three decades. Errol Morris, who directed the film version of A Brief
History of Time in 1990, says that Hawking gave him a copy of his voicebox
software so he could record Hawking’s voice without the scientist’s
presence. “Theoretically, I could have Hawking saying anything. There’s
something quite absurd about that,” Morris says. “You just type in a sentence
and you record it, and you put it in the movie.” At some point, Morris
tweaked one of Hawking’s statements and the scientist noticed the alteration
immediately. “He said, ‘You changed that,’” Morris recalls. “And then he
said, ‘But I like it better.’”26
Unlike a regular actor, Stephen Hawking could never flub his lines; if they
were entered correctly into the computer, they’d come out as written every
single time. And because Hawking didn’t move his lips when he talked, a
director with the full use of the physicist’s voicebox could superimpose
Hawking’s speech on any image of the physicist sitting in his wheelchair, and
it would seem like Hawking himself was intoning the words. No matter if
Hawking himself hadn’t composed, or even heard, what he was telling the
audience. When Hawking outsourced his voice to a film production, he
granted a director almost unheard-of control, an incredible money-saving
boon. Shoot a few scenes of the physicist sitting in his chair (typically slow-
moving spiral shots in a wood-paneled hall) and a few close-ups of his eye
moving about or his head settled uneasily on his shoulder, and that’s all that
would be necessary for a production. The film editors could mix and match
any of those images with whatever words were passed through the scientist’s
speech synthesizer.x
In 2016, Hawking put his name on an oddball series that aired on public
television stations in the United States. Part reality show, part science
documentary, Stephen Hawking’s Genius asked contestants to tackle scripted
challenges that illustrated scientific principles. (In one show, for example, a
team was asked to melt a bucket of ice without an obvious source of heat;
they wound up bending a metal bar back and forth to convert mechanical
energy into heat energy.) The series consisted of six hour-long episodes. Yet
the director needed less than four minutes of Hawking footage to carry
through the entire season. The film editors used the same tiny library of shots
over and over—cut and spliced in different places, digitally colored or
otherwise altered in subtle ways, or even time-reversed to make them seem
distinct. Only the most careful of observers would be able to tell that the
exact same video footage showed Hawking talking about chemistry in one
episode, evolution in another, and the expansion of the universe in a third.27
“You know, I would joke to him that he was the first non-talking talking
head,” says Errol Morris.28
To a slow, ambling beat, the electric guitar laments, wailing a gentle, feline
wail, and then fades. Out of nowhere, a voice, a robotic voice, takes the lead.
“Speech has enabled the communication of ideas, enabling human beings to
work together to build the impossible,” Hawking says, keyboard and guitar
rising higher and higher in pitch as he continues his monologue. “Mankind’s
greatest achievements have come about by talking.”
It wasn’t Hawking’s first appearance on a Pink Floyd album; two decades
earlier, the physicist had provided vocals for the band’s “Keep Talking.” In
2014, Floyd decided to release its first studio album in twenty years, and it,
too, included a Hawking-narrated track, “Talkin’ Hawking.” It was just as
trippy, and Hawking’s voice just as incongruous, as the first song.
Hawking was on his way to becoming a regular fixture in the world of
rock ’n’ roll. In 2015, he went on the road with the band U2—albeit virtually
—on their iNNOCENCE + eXPERIENCE tour. At each show, fans were
treated with a video in which Hawking declared, “One planet. One human
race. We are not the same, but we are one.” This line elicited cheers, but not
as much as when he told the audience, “We give our elected officials their
power, and we can take it away.”29
By the time Hawking received payment for the U2 video, his foundation
was up and running. At a gala dinner at the Royal Institution in London in the
fall of 2015, Hawking and a variety of celebrities—including Eddie
Redmayne—gathered to launch the physicist’s new enterprise. Dedicated to
promoting cosmology and helping people with ALS, the foundation began
giving small grants to send young Britons to space camp, creating science
tunes to help teach young children, and funding research into the very early
universe. The foundation—overseen by Hawking’s sister Mary; his friend
Kip Thorne; and his Cambridge colleague Malcolm Perry—quickly started
raising money.
In its first financial statement, the Stephen Hawking Foundation declared
that it had earned £26,000 from a number of sources: rights from the U2
video, from marketing a set of postage stamps issued on the Isle of Man (one
of the stamps featured the left side of Hawking’s visage, perfectly mirroring
the right side of Einstein’s face immortalized on a neighboring stamp), and
from the sale of scented candles. It wasn’t a huge amount; straight-up
donations from outsiders netted roughly twice as much.30
Given how little was flowing into the foundation’s coffers, it’s likely that
Hawking himself was also receiving a relative pittance from trading on his
name. Despite his agent’s attempts to find more sources of funding, Hawking
had decided to make a change.
“He had a meeting with me,” says Al Zuckerman, Hawking’s longtime
agent. Zuckerman had started working with Hawking in the early 1980s, just
as the physicist decided he wanted to write a popular book—the one that
became A Brief History of Time. “I helped to make a lot of money [for him],”
Zuckerman says with a grin. “Or should I say he helped me make a lot of
money.” Despite working with Hawking for so long, he had never signed the
physicist to an exclusive contract, and he didn’t handle all of the physicist’s
moneymaking ventures. “He had an office with a succession of different
people, and they were besieged with requests for his appearances—some
universities, and some conferences—and they were really in a position to
make money for him,” Zuckerman says. “But the people he had running [his
office] were not equipped very well to do that, and so most of those requests
did not come to me.” It wasn’t the most efficient arrangement, but it seemed
to work for both Hawking and Zuckerman. So the meeting came as a
complete surprise.31
“So he met me and [his lawyer], who decides to fire me. Which she did,”
Zuckerman says. “I think that he wanted more money than he was earning.
And I set myself up to do a lot of the search for ways to get more income. But
when push came to shove, he opted to go with this Brit guy.” (This Brit guy
being Robert Kirby, an agent who represented Hawking in his final two years
and, at the time this book is being written, acts as agent to the physicist’s
estate.)
Roughly three years after that meeting, Zuckerman still seems a bit
stunned. “He opted to go with Kirby. I have no idea why. I did a lot of work
on trying to figure out how to find money for him, and I don’t know what I
said,” Zuckerman explains. “I think I made a mistake, though, when I was
there, because I was really mostly talking to his lawyer, who, you know,
could click with and understand his speech and respond, whereas he was
turned away and looking at a computer all the time. I should have realized
that I should not have been looking at her but looking at him. But, uh,”
Zuckerman sighs. “Fortunately, you know, my life, my livelihood no longer
depends on…” His voice trails off.
Stephen Hawking seldom wagers much money. It’s a good thing, since he
usually loses. Most of the time, it isn’t by design.
This time, it looked like he had backed a winner. On March 17, 2014, a
team of American physicists announced that they’d detected the subtle signal
of gravitational waves—not from black holes crashing together, but from the
Big Bang itself. When Hawking spoke on BBC radio the next day, he told
listeners, “Yesterday a team from Harvard announced they had detected
gravitational waves from the very early universe.” Then, in an announcement
that would be stunningly bathetic were it coming from anyone other than
Stephen Hawking, he continued, “It also means I win a bet with Neil Turok,
director of the Perimeter Institute in Canada.” With that, Hawking’s wager
became part of the ongoing story. And a big story it was, as it seemed to be a
snapshot of the very first moments after creation.32
A cataclysmic cosmic event, like the spiral-and-crash of two black holes,
causes ripples in the fabric of spacetime, so it should come as no surprise
that the biggest cataclysm of them all—the Big Bang and the early rapid
inflation of the universe—sent shudders throughout the spacetime fabric of
the cosmos. Those gravitational waves are so stretched and attenuated by the
passage of time and the expansion of the universe that we can’t detect them
directly. Even LIGO’s ultra-sensitive detectors are too weak and small to
spot these waves. But those gravitational waves left their mark in the
heavens.
Many billions of light-years away, boxing us in in every direction, we’re
surrounded by walls of light. These walls are invisible to the human eye;
their ancient light has been so stretched by the expansion of the universe that
only special microwave detectors can spot them. But they’re there
nonetheless, everywhere in the sky at once. They are the afterglow of the Big
Bang, the light from a moment about four hundred thousand years after the
birth of the universe when the glowing-hot clouds of gas that filled the
cosmos cooled enough to suddenly become transparent, liberating the light
that had been trapped within. That primordial light, now known as the cosmic
microwave background (CMB), is ubiquitous—no matter which way you
point your telescope, it’s there. And it’s the most distant object we can see.
Across those walls lie the very young universe, the cosmos when it was
younger than four hundred thousand years old.xi From beyond the CMB, from
beyond those walls, light simply can’t reach us.
That’s where gravitational waves come in. Even though we can’t see light
from the very early universe, gravitational waves aren’t blocked by those
walls. What’s more, the rippling fabric of spacetime affects how those walls
look—the gravitational radiation stretched and squished the primordial
clouds of gas, and affects the nature of the CMB. Specifically, scientists were
looking for telltale signs of gravitational waves in the polarization of the
cosmic microwave background radiation.xii And on March 17, 2014, when
the researchers associated with a sensitive Antarctic microwave telescope,
BICEP2, announced that they had found those signs, the physics community
came alive with excitement. Much of the community, anyway.
“This is huge, as big as it gets,” Marc Kamionkowski, a theoretical
physicist, told the New York Times for its front-page story. “This is a signal
from the very earliest universe, sending a telegram encoded in gravitational
waves.” Newspapers and press outlets were full of enthusiastic quotations
from theorists who had studied the early universe, including Alan Guth (who
came up with the theory of inflation describing the rapid expansion of the
universe) and Andrei Linde (who helped devise the version of inflationary
theory that is most popular today). Linde had even celebrated the discovery
with a bottle of champagne brought over by a BICEP2 scientist. Physicists
were even beginning to mutter about a crowning glory for the discovery; the
chair of the Harvard Astronomy Department, theorist Avi Loeb, told the
Times that if the results held up, “it’s worth a Nobel.”33
When Hawking went on BBC the following day, he was thrilled about the
news; BICEP2’s results implied not just that they had detected a signal from
the very earliest moments of the Big Bang, but also seemed to show that Alan
Guth’s and Andrei Linde’s description of the early universe—the theory of
inflation—was fundamentally correct. That’s what Hawking’s bet had been
about.xiii
Hawking’s former colleague, Neil Turok, along with several other
cosmologists, such as Princeton’s Paul Steinhardt and the University of
Pennsylvania’s Burt Ovrut, had been working on an alternative theory that
did away with the rapid expansion of the early universe.xiv In their
formulation, there wouldn’t be any primordial gravitational waves rattling
around the early universe, hence no imprint of gravity upon those hot clouds
of primordial gas. Hawking, however, was all-in on inflation; not only was
he close friends with Linde—and antagonistic toward Steinhardt—but
Hawking had played a major role as midwife for the birth of the theory of
inflation (see Chapter 11). So, a true believer in inflation, Hawking had bet
Turok $200 that standard inflationary theory was correct—and that there had
to be primordial gravitational waves. Or, just as importantly, Turok, Ovrut,
and most especially, Steinhardt, were wrong.
Based upon the buzz in the theoretical community, it seemed that Turok
would have to pay up. For once, Hawking had won a bet. However, Turok
wasn’t so sure. “I have some reasons for doubt about the new experiment and
its results,” he said later that day. “It’s not entirely convincing to me that they
have clearly seen what they have claimed to have seen.” This wasn’t just
sour grapes. Nobody had yet had time to flyspeck the data, to assess for
themselves whether the BICEP2 team had made a great discovery or just an
embarrassing mistake.34
From a theorist’s point of view, the BICEP2 observations were thrilling,
because they suddenly gave direct support for an important segment of
cosmology’s theoretical scaffolding—inflationary theory. From an
experimentalist’s point of view, it was an exciting but technically challenging
observation in which a lot of things could go wrong. Seeing a pattern of
polarized light in the sky might well be a signal from the early universe; it
might also be light bouncing off dust clouds. Telling the difference between
the two is not so easy. And there was a concrete reason to worry: an
expensive microwave-detecting spacecraft, Planck, hadn’t seen anything of
note. If, in fact, BICEP2 was right, Planck should have also spotted the
signal—and there was no good explanation about why the Antarctic
telescope was seeing something that the spacecraft didn’t.
Even if Hawking had enough time to read the BICEP2 paper carefully, as
a die-hard theorist, he didn’t have the experimental chops to find potential
flaws in the analysis. Neither did Turok, for that matter, but his skepticism
turned out to be warranted. Within weeks, eminent experimentalists began to
poke holes in the BICEP2 team’s analysis. It was a rather technical argument:
when the BICEP2 team tried to subtract out the effects of dust using a
mathematical model, they had done it in the wrong way. And time proved the
skeptics right. More evidence—observations using multiple frequencies of
microwaves rather than just one—showed that the polarized light came from
dust, not from primordial gas clouds. By early 2015, the BICEP2 researchers
had withdrawn their claim.
Hawking hadn’t won the bet after all. But neither had he lost. That’s more
than can be said about his other bets, including the ones closest to his
wheelhouse: the physics of black holes.
Perhaps the most striking property of a black hole is how featureless it is.
The event horizon that shrouds the center of the black hole is completely
opaque—nothing, no light, no matter, no energy, no information at all—can
cross that invisible boundary and escape a black hole. This means that an
observer outside the event horizon can tell nothing at all about what’s on the
other side of the horizon. The information blackout is almost total.
That’s one of the most important conclusions physicists reached during the
Golden Age of Black Holes. Hawking and others argued that because no
information can escape the event horizon, an observer coming across a black
hole could learn almost nothing about it: its age, its composition, what kind
of matter it had swallowed since it was born—these were unfathomable
mysteries hidden behind the veil of the event horizon. In fact, the physicists
concluded, there were only three things that one could tell about a black hole
from the outside: its mass, its charge, and how fast it spins. Nothing beside
remains.
A black hole is almost totally devoid of distinguishing characteristics, a
dictum that became known as the no-hair theorem.xv However, starting in the
mid-1970s, Stephen Hawking came to realize that this featurelessness, which
is required by the theory of relativity, causes immense troubles for theorists.
A different set of laws—the laws of quantum mechanics—says just as firmly
that black holes simply cannot be that featureless. Quantum theory says that
somehow, the history of a black hole, the information about what it was made
of and what it had swallowed, had to be preserved. And Hawking’s own
research implied that even a black hole can’t hide such information from
view forever; it either had to be destroyed (violating quantum theory) or
emerge somehow (violating relativity). This contradiction—the information
paradox—quickly became one of the biggest scientific puzzles of the day. It
went to the heart of the nearly century-old fissure at the heart of modern
physics, the mutual incompatibility of quantum theory and relativity, the two
great physical frameworks of the twentieth century. Solve it, and it was
possible that the solution would reveal the ultimate answer. It might show the
way to an overarching, unifying theory that incorporated both relativity and
quantum theory, something that described all the matter and energy and forces
in the universe on all scales, large and small. Hawking had been groping for
a solution to the information paradox ever since he described it, and toward
the end of his life, he believed that the solution to the paradox was likely
sitting right on top of a black hole’s event horizon.
In 2014, when Hawking announced that “there are no black holes,” it
wasn’t really a declaration of the nonexistence of the objects he had studied
his whole career, but a salvo in his attempt to resolve the paradox by
reexamining the physics of the event horizon. Though Hawking rejected the
recent argument floated by Joe Polchinski and other physicists—that there
was a firewall at the event horizon that incinerated infalling matter (and
prevented information from getting lost)—he apparently agreed that
something different was happening at the horizon, something that could
resolve the paradox. Nobody understood quite what Hawking was driving at,
not even Polchinski, who dubbed Hawking’s nascent idea chaos walls. But
the scientific community awaited further information.
Despite the headlines, Hawking never provided any more information. He
soon abandoned the concept of chaos walls in favor of an idea he found much
more exciting. A few months after publishing the chaos-wall paper, Hawking
visited a ranch owned by a billionaire Texas oilman—George P. Mitchell—
for a physics retreat. Mitchell was a major admirer of Hawking’s; he had not
only endowed a chair at Texas A&M in Hawking’s name, but named a brand-
new auditorium on campus after Hawking. Periodically, Mitchell would
invite Hawking and some of his colleagues to gather and discuss the
mysteries of the universe. This time, Andy Strominger, a physicist at Harvard
who had known Hawking since the early 1980s, was there, and Strominger
had some new results.35
Strominger had been working on a branch of general relativity that had
been discovered—and abandoned—in the 1960s, and he was increasingly
convinced that the ideas he was coming up with could help physicists
understand what was going on at a black hole’s event horizon. “And Stephen
was very excited by this. I gave a seminar in the afternoon, and we stayed up
until one in the morning,” Strominger says. “He wrote, I haven’t been this
excited since… I feel like I felt when I discovered the area law.xvi And also
his nurses said that they hadn’t seen him that excited in forever.… He became
very energized.” And Strominger started working with Hawking and fellow
Cambridge physicist Malcolm Perry to refine the idea. By late 2015,
Hawking was publicly hinting that he, Strominger, and Perry were thinking
about something new and exciting, promising that he was working on a “full
treatment” of the concept.36
Collaboration with Hawking was always a bit challenging. “Malcolm and
I would be writing at the board and getting cues from Stephen,” Strominger
says. “Stephen would sit there typing, and there was always a few minutes’
delay.” But the two were used to the awkward flow of a conversation with
Hawking. More problematic was the difficulty of getting the three physicists
in the same place at the same time. After several visits to the United
Kingdom, Strominger invited Hawking and Perry to visit Boston. “There was
always a money problem,” says Strominger. “I mean, at some point some of
his rich friends were offering their jets to fly him around, but he couldn’t go
on their jets anymore. He could only go on an ambulance jet, and not only
that, his doctor insisted on one Swiss company, and it got very expensive…
so I had to somehow find the money for him to come.” Eventually the money
came through—the visit would cost hundreds of thousands of dollars—thanks
to one of those rich friends.37
But even with the money, the trip wasn’t possible unless Hawking’s
doctor signed off on the trip. This time, the doctor reportedly said no.
Hawking wrote a letter asking him to relent. “He wrote about how excited he
was about the research and that it was potentially going to be a Nobel Prize,”
Strominger recalls. “I think his doctor was not very sympathetic to having
him go off and meet celebrities, but when it came to his legacy and his
science… and I think everybody could see that Stephen was invigorated by
our new enterprise in a way that he hadn’t been for quite some years.” And
so in April 2016, Hawking prevailed; his doctor let him fly to Boston, where
the three continued their work.
Despite all the difficulties, Hawking, Strominger, and Perry had a very
productive collaboration. They soon delivered a series of papers arguing that
the “no-hair theorem” wasn’t quite right: black holes were, in fact, covered
with “soft hair.”
The idea is that when a charged particle falls toward the event horizon of
a black hole, the rules of electromagnetism dictate that a bizarre, energyless
particle of light—a “soft” photon—is born right at the black hole’s event
horizon. Such a photon not only stores information about the particle that
created it, it will eventually escape its precarious position right on top of the
event horizon. That is, soft photons record information about matter that has
crossed the event horizon; at some point in the future, this information will be
visible to an outside observer. A black hole isn’t truly a featureless object,
Hawking’s argument went. It is not a void whose history is totally lost to the
bare and boundless sands of time. It is a void with a recorded history, a
history stored on soft photons at the event horizon. As no two black holes
share the exact same history, their collections of soft photons must be
different. Each one has its own unique head of “soft hair.”
Hawking seemed convinced that, at long last, he had solved the problem
of what happens to particles when they fall into a black hole—the question at
the core of his scientific legacy. It would be his crowning achievement.
His collaborators were not so sure. “Stephen wasn’t afraid of simplifying
our work to get to the essence of it and to convey the excitement, but
sometimes he said some things that were, you know, ‘We’ve solved the black
hole [problem],’” Strominger says. “Well, even worse, he’ll often replace
the ‘we’ with ‘I.’”38 xvii
Physicists outside the team were even less enthusiastic about the soft-hair
idea. Hawking’s former PhD student Raphael Bousso and a colleague argued
that a “bad choice” in the team’s mathematics had led to the wrong
conclusion. “In other words, the soft hair is a wig,” he wrote. Its roots didn’t
reach down beyond the event horizon—it was not connected to the innards of
the black hole. The idea that this hair was preserving information that fell
past the event horizon was merely an illusion. That is, soft hair bore “no
relevance to the black hole information paradox.” Another former student,
Marika Taylor, agrees. “Stephen was so smart. He had to know that, maybe
this kind of gave a contribution, but it was not solving fundamental puzzles,”
she says. “There’s many different ways to see that this was not going to be
the answer. And Stephen had to know that. But… he likes the adulation when
he does something big.”39
Even though soft hair was not the breakthrough Hawking seemed to claim
it was, for the first time in more than a decade scientists were actively
grappling with one of Hawking’s papers—he had sparked a serious debate
among his fellow scientists.
Once again, Stephen Hawking was at the center of a controversy; once
again, the battleground was the event horizon, the boundary between the
universe we know and the darkest unknowns known to humanity. However,
this time, for the very first time, the experimentalists had begun to venture
there as well.
Every year, rumors seem to fly faster than ever before. On social media, a
quiet blurt can spread around the world at the speed of light. Thousands can
hear it and amplify it within a matter of minutes. This is true even in the
world of science. And on September 25, 2015, a stray comment on Twitter
by a cosmologist in Arizona set off a bout of fevered speculation, leading
once more to whispers about Nobel Prizes. This time, though, the outcome
would be happier than it had been for the BICEP2 team.
“Rumor of a gravitational wave detection at LIGO detector,” cosmologist
Lawrence Krauss tweeted late that Friday afternoon. “Amazing if true. Will
post details if it survives.”40
It had been just a few weeks since the LIGO detector had finished a long
upgrade and turned on once again. Physicists working on the project had been
waiting nervously—LIGO had been unexpectedly silent since it had first
turned on in 2002. Had its builders finally made good on the promises they’d
made decades (and hundreds of millions of dollars) ago? The LIGO team
said nothing, and after a brief flurry of blog posts and short stories in the
specialist press, the buzz quieted down. For almost three months.
Then, Krauss struck again. “My earlier rumor about LIGO has been
confirmed by independent sources. Stay tuned! Gravitational waves may
have been discovered!! Exciting.” This time, the rumors wouldn’t go away.
For almost a month, there was a steady drumbeat of speculation, which
reached a fever pitch on February 8, 2016, when the collaboration declared a
press conference three days hence. The team was coy about the precise
nature of the upcoming announcement—merely saying they were going to
provide an update, a “status report on the effort to detect gravitational
waves.” Even though it was the worst-kept secret in physics, the LIGO team
itself was being extremely careful not to let things slip before the formal
announcement. They almost succeeded.41
Sixteen minutes before the press conference was set to begin, a NASA
astronomer used Twitter to post a picture of a cake decorated with a picture
of two black holes spiraling into each other. “Here’s to the first direct
detection of gravitational waves!” it announced, in green icing. The rumors
were true. A pastry had let the Einsteinian cat out of the bag.42
A few minutes later, Caltech’s Kip Thorne was explaining the discovery
to a crowd of journalists. “There was one regime in which general relativity
had never been tested,” said Thorne. For the first time, LIGO was giving a
direct view of a place where the fabric of spacetime is not just extremely
warped, but changing very rapidly—the conditions right near the edge of a
black hole as it swallows a large lump of matter. “We have never had any
tests in that regime,” Thorne continued. “This observation tests that regime
beautifully, very strongly, and Einstein comes out with beaming success.”43
It was a success that Hawking could hardly have dreamt of when he was
starting his career. Back then, and on through the Golden Age of Black Holes
in the late 1960s and early 1970s, physicists couldn’t even prove that black
holes existed. The collapsed stars were the theoretical byproduct of the rules
of general relativity; by following those laws to their logical consequences,
Hawking and Thorne and numerous other colleagues were able to describe
black holes in great detail. However, that wasn’t the same as actually being
able to point to a black hole in the sky. Finding a black hole—an object that
absorbs light that ventures too close—was an incredibly difficult task. As
Hawking put it, trying to find one was akin to “looking for a black cat in a
coal cellar.”44
By the time Hawking was starting his black-hole research, astronomers
had seen a strange object in the constellation Cygnus—a mysterious mote in
the sky that shone brightly with X-rays. Scientists suspected that it was a
dead sun, and if it was, it was far too heavy to be any of the other kinds of
collapsed stars that theorists knew about. It could just possibly be a black
hole. Most astrophysicists came to the conclusion that it was a black hole,
but the evidence was indirect. Indeed, Hawking’s most famous wager—his
1975 bet with Thorne—was about whether this weird object in Cygnus was,
in fact, a black hole. Hawking only conceded the bet in 1990.
Over the years, astronomers built up better and better evidence for the
existence of black holes. Using X-ray detectors, they spotted more objects
like the one in Cygnus. Using infrared and visual telescopes, they peered at
the centers of distant galaxies looking for signs of massive black holes
swallowing matter. (Astrophysicists now think that pretty much every galaxy
has a black hole at its center.) They watched as stars at the center of our
galaxy wheel around a massive invisible object. Astronomers are very
certain that black holes exist, but using telescopes that detect different kinds
of light—X-rays, ultraviolet light, visual light, infrared light, microwaves,
radio waves—it’s exceedingly hard to spot an object that doesn’t let light
escape, much less to probe its properties. The theorists with their equations
were seeing far deeper into the abyss than the experimentalists with their
instruments could possibly hope to go.
That’s what LIGO had suddenly changed in September 2015. A brief
shudder of spacetime heralded the cataclysmic collision of two massive
black holes. And every few weeks, LIGO was detecting yet another such
“merger”—one in October, another in December. Now, by using gravitational
waves rather than light waves, experimentalists were not just detecting black
holes, but beginning to gather data about the region not far from a black
hole’s point of no return: its event horizon.45
As two black holes spiral in toward each other, they orbit faster and
faster and draw closer and closer, emitting gravitational waves all the while,
setting spacetime all aquiver. In the final milliseconds before the collisions,
the black holes’ event horizons come into ever closer proximity. The
gravitational waves that the holes emit in the very last split seconds before
that colossal wreck, the last, violent tremors before all is silent once more,
are signals from near those event horizons. Experimentalists were beginning
to venture right up to the edge of the abyss—just at the border where they
might be able to verify some key predictions that Hawking had made about
black holes in the 1970s. If they could, Hawking would—after half a century
of waiting—like Einstein, have an experiment demonstrating that his ideas
were correct.
“He was very excited,” Thorne says. “He wanted to know how well we
can measure the masses and spins of the black holes in order to test his area
theorem.” Unfortunately, for technical reasons, LIGO isn’t precise enough to
provide a stringent test of any of Hawking’s ideas directly: not his work from
the 1970s, much less his recent fight over soft hair and firewalls. Even so, it
was a stunning result. An observatory was for the first time scoping out the
battlefield where Hawking had spent so much of his career. He certainly saw
it as a victory. “Along with confirming Einstein’s beautiful theory, the
detections agree with predictions that I and other scientists have made about
black holes,” Hawking declared in late 2016. “The ripples of their work will
flow through the field of astrophysics for many years to come.”46
The following October, Barry Barish, Rai Weiss, and Kip Thorne won the
least surprising Nobel in recent memory, “for decisive contributions to the
LIGO detector and the observation of gravitational waves.” At the moment of
the announcement, the Nobel committee published copious information about
the laureates’ achievements. There was a press release announcing the basis
for the prize; a popular account intended to make the science accessible to
the public; and a dense, eighteen-page background going deep into the
winners’ work and its importance. In all of that material, with all of its
dozens of references and kudos and historical explanations, one name was
entirely missing.
Nowhere acknowledged at all: Stephen Hawking.
Footnotes
i As physicist John Wheeler famously put it, “Spacetime tells matter how to
move; matter tells spacetime how to curve.”
ii In retrospect, it’s not at all clear that Eddington’s measurements were
sufficient to prove Einstein correct, but that mattered little at the time.
iii As with Eddington’s eclipse observations, these measurements, by
astronomer Walter Adams, were in fact too error ridden to be good tests of
general relativity. Later, however, more accurate observations verified
gravitatgional redshift (and gravitational lensing) with high precision.
iv But again, the “fagbric” is really a four-dimensional manifold rather than a
two-dimensional sheet, and includes time as well as space. In addition, the
ripples are only created under certain circumstances, so the analogy is
misleading in some respects.
v Except, technically, for tidal forces: a stretching pull that eventually turns
the astronaut into spaghetti as he or she falls toward the singularity. For
sufficiently big black holes, that tidal force should be negligible at the event
horizon.
vi The degree to which Hawking and his estate were and are in control of
how third parties use his synthetic voice is an interesting legal question—it’s
a computer program, created by others, and can thus be used by others, yet it
is associated with him just as surely as his face. While I have found
trademarks associated with Hawking’s name, I have not found any evidence
of copyright on the voice or even an indication whether such a copyright
could be possible.
vii Font size is apparently a big deal in Hollywood circles.
viii Even with a box-office success like this, the author won’t necessarily
make a huge amount of money. An SEC filing from 2006 revealed the terms
of Jane’s film-rights contract for Music to Move the Stars, in which she was
paid an option price of $2,000 and promised 2.5 percent of the production
budget and the same share of the net profits of the film should it ever be
made. If her Travelling to Infinity contract was similar, she probably would
have made at most a few hundred thousand dollars from the deal.
ix At his peak, Hawking could only speak at about 15 words per minute;
more typical was 3 words per minute or below. For comparison, a good rule
of thumb for English speakers is about 120 words per minute.
x Even so, Hawking could be a difficult actor. One director of a 2005
documentary told sociologist Hélène Mialet how Hawking showed up
extremely late to a shoot for a BBC documentary, then left early in a fit of
pique. Later, attempting to get the Hawking voiceover, the director looked
“on the Internet to try to find a kind of Stephen Hawking sound-alike, but we
couldn’t actually find one.” Eventually, they convinced the physicist to
“read” out the script that they had prepared. Hélène Mialet, Hawking
Incorporated: Stephen Hawking and the Anthropology of the Knowing
Subject (Chicago: University of Chicago Press, 2012), 96–98.
xi Because light travels at a finite speed, looking at a distant object is like
looking back in time. It takes about eight minutes for light to travel from the
sun to the Earth, so when you look up at the sky and see the sun, you’re seeing
light that emerged from the surface of the sun roughly eight minutes ago. Light
from the nearest galaxy, Andromeda, is more than two million years old.
xii Light waves have a certain degree of directionality. They can be
“waving” vertically, or horizontally, or in circles—think of all the different
ways two children can wiggle a jump rope. By examining patterns of this
directionality, this polarization in the CMB light, scientists could, in theory,
figure out how gravitational waves were affecting the clouds of gas in the
very early universe. For more information, see my book Alpha and Omega,
209 ff.
xiii Paul Steinhardt and his graduate student, Andy Albrecht, also deserve
credit, which Hawking was loath to give.
xiv For more on this idea, known as the ekpyrotic or cyclic theory, see my
Alpha and Omega, 196 ff.
xv The term came to be when black-hole physicist Jacob Bekenstein, then a
graduate student at Princeton, struck by a black hole’s cue-ball-like
featurelessness, exclaimed that “a black hole has no hair.” His adviser, John
Archibald Wheeler, popularized the phrase—but not everybody was happy
with it. Physicist Richard Feynman “thought that was an obscene phrase,”
Wheeler once told an interviewer. “He didn’t want to use it.” Interview with
John Wheeler, Part 84, “Feynman and Jacob Bekenstein,” available at Web
of Stories, www.webofstories.com/play/john.wheeler/84.
xvi One of Hawking’s key insights during the Golden Age of Black Holes.
xvii Strominger emphasized that he bore no resentment. “As I said to my
friend, you write a paper with Stephen, you get ten thousand times as much
attention as if you write it without Stephen. If I did 99 percent of the work,
I’m still coming out ahead by a factor of a thousand,” he said, laughing. “If
you look at Stephen’s stuff, he almost never gives credit to any of his
collaborators when he’s interviewed. I think he’s given more credit to
Malcolm and me than I’ve ever seen him give anybody. Sometimes he sort of
suggests that we’re working for him, but you know, I really don’t feel
cheated. I’ve had plenty of appreciation so I’m not about to complain, but it’s
a fact that he does this.”
CHAPTER 3
MODELS (2012–2014)
The stadium was bathed in an otherworldly blue glow, a color so vivid and
pure that it seemed unnatural. At center, dwarfed by an enormous model of
the moon perched on a tower high above, sat a tiny figure in a wheelchair. He
was almost invisible, an immobile speck of dull brown in the middle of a
gaudy, twinkling whirl of motion—a spectacle that had suddenly paused in
anticipation. A hush fell over the crowd.
“Ever since the dawn of civilization, people have craved for an
understanding of the underlying order of the world: why it is as it is and why
it exists at all,” Hawking’s mechanical voice rang out over the loudspeaker.
The physicist sat perfectly still as his speech synthesizer continued. “But
even if we do find a complete theory of everything, it is just a set of rules and
equations. What is it that breathes fire into the equations and makes a
universe for them to describe?”1
As Hawking fell silent, a giant, glowing, smoking ball—resembling an
atom with electrons circling about, or perhaps an armillary sphere
representing the positions of the stars in the sky—descended from the
heavens into a blue-lit void. Fireworks erupted in all directions as
ponderous music poured from the loudspeakers. The 2012 Paralympic
Games had begun with a Big Bang.
London’s brand-new Olympic Stadium was nearly full to capacity. At
over sixty thousand people, it was the largest assembly Hawking had ever
addressed. However, even though he was in front of the biggest audience of
his life—and was part of a production that had expended more money and
was reaching more viewers than any of his prior appearances—Hawking had
not gone out of his way to do anything special for the occasion. Completely
unfazed by the pyrotechnics around him, he spoke as he always did, his
voicebox pronouncing the same words that it had pronounced numerous times
before.2
Hawking was often touted as one of the finest communicators of science
in the world. However, he had to go through extraordinary lengths to make
himself understood by the people around him. After losing his voice in the
mid-1980s, he relied almost totally on the computer embedded in his
wheelchair to speak and to write his books and speeches. Using a clicker in
his right hand (and then, as his disease progressed, a sensor near his cheek),
he navigated through a menu of options to select the words he wanted to utter,
one by one. Only after Hawking had painstakingly assembled a string of
words could he then instruct the voicebox to pronounce them aloud. At his
peak, Hawking could get about fifteen words per minute out of the machine.
But as his muscle control declined, his communication speed dropped as
well.
“When I first met him, he was speaking in real time—maybe he spoke a
little more slowly than the average person,” Andy Strominger says. “Then,
twenty years ago, you would talk about words per minute, and then it went to
minutes per word, and then, you know, it just got slower and slower until it
just sort of stopped.”3
It was often easier for Hawking simply to call up an old sentence—one he
had sent to his computer previously that was still stored in its memory bank
—than to compose the same or a nearly identical one from scratch. As a
consequence, his public utterances were often pastiches, bits and pieces of
previous writings and statements fitted together like a mosaic. Even before
Hawking’s mechanical voice made this habit a labor-saving device, he
seemed to delight in feeding the same bon mots to journalists over and over
again. As one journalist recalled in the mid-1980s, shortly before Hawking’s
tracheostomy:
Not that it mattered to his audience at the Paralympic Games, but the other
parts of Hawking’s opening were also a quarter century old. “Ever since the
dawn of civilization, people… have craved an understanding of the
underlying order of the world,” comes from his most famous book, A Brief
History of Time, as does “Even if there is only one possible unified theory, it
is just a set of rules and equations. What is it that breathes fire into the
equations and makes a universe for them to describe?” Even Hawking’s now-
famous exhortation at the Games to “look up at the stars, not down at your
feet” was repurposed from a few years prior.5
But there was one section of Hawking’s address that was entirely novel,
something that Hawking had never said before, at least in public. As his
speech drew to a close, Hawking told the gathered crowd that the recent
discovery of the Higgs boson at CERN was a triumph of a lifetime. “It will
change our perception of the world and has the potential to offer insights into
a complete theory of everything,” he declared through his electronic
voicebox as two glowing-visored DJs gesticulated wildly behind him.
Hawking’s voice faded and the sound system began to thrum with electronic
dance music. Acrobats spun and gyrated, twirling red umbrellas. Actor Ian
McKellen waved a placard to the beat, and backup dancers in wheelchairs
pumped their fists in the air. Only about a month earlier, particle physicists
had announced that the Higgs had been found, and already it had been
repurposed for popular entertainment. Hawking sat, immobile, center stage, a
striking contrast to the commotion surrounding him.
Hawking always enjoyed being at center stage, and in the last years of his
life, he was used to being there. At least in popular perception, he inhabited
the very center of the scientific universe. Like Newton and Einstein, Hawking
was the shining star of physics, obscuring all others by his sheer brilliance.
Other pretenders to the throne would sometimes arise, other scientists
who would be talked about almost as much as Hawking, at least for a short
time. In 2012, for example, a massive particle accelerator on the French-
Swiss border put Scottish physicist Peter Higgs in the headlines. For that
multibillion-dollar machine was finally bringing to a close a forty-year-old
quest to find a particle that Higgs (and others) had predicted: the Higgs
boson. Or, as one Nobel laureate had dubbed it, the “God particle.”
Unlike most of his peers, Hawking was a Higgs atheist; he didn’t believe
in the God particle, at least not until he was forced to. However, even had
Hawking not had some unorthodox scientific views about particle physics, it
was almost inevitable that he would get into a fight with Peter Higgs at
Higgs’ moment of greatest triumph.
Hawking wasn’t afraid of controversy. His ideas were often
unconventional, even heretical; even as he entered his scientific twilight, he
and his collaborators kept trying to generate new theories to wield in
scientific battles. (Sometimes Hawking’s battles could turn nasty; he had
gravely offended colleagues on occasion.) As a celebrity, however, Hawking
had to deal with a whole other level of scandal, including tabloid gossip
about his sex life. Of all the controversies that Hawking got involved with
over the years, the one over the Higgs boson is a minor one. Yet for
Hawking, it was the one that is most clearly born from sadness—sorrow that
the universe wasn’t turning out to be as interesting as the model Hawking had
painstakingly built up in his mind.
It is almost as if the Higgs boson—the discovery that won Peter Higgs his
Nobel—comes from a different universe than the one described by Einstein’s
general theory of relativity.
Einstein’s cosmos is filled with a smooth, gently rippling manifold that
stretches over almost unfathomable expanses of space and time. The force of
gravity comes from the way matter and energy cause that manifold to curve.
And while the curvature can be enormous on a large scale, in tiny,
submicroscopic regions it becomes almost negligible. With the sole
exception of a black hole—where the curvature is so extreme that the
smoothness of the manifold breaks down—spacetime on the smallest scale is
a continuous, quiescent, almost boring place.
The universe of the Higgs is not smooth manifold, but a frothy,
discontinuous substrate that is ever more violent the smaller the scale. It is
not a realm where forces are the result of how spacetime curves; instead,
forces are carried by undetectable subatomic particles that pop into
existence, carom off each other, and disappear once again into the void from
whence they sprung. The ever-churning sea of the particle physicist looks
almost nothing like the rubber sheet of relativity. This disconnect is the
biggest problem of modern physics, and one that Hawking spent much of his
career trying to understand.
Quantum theory describes how matter behaves on very small scales: how
molecules and atoms and subatomic particles move and interact with each
other. Embedded into the laws of quantum theory are some rules that are
responsible for a froth, a churning, random activity of particles on the very
tiniest segments of space and time. Like relativity, quantum theory was born
at the turn of the century, and as with relativity, Einstein played a central role
in its creation. But Einstein himself was repulsed by the randomness and the
discontinuous nature of the universe of quantum theory. “God does not play
dice with the universe,” he declared. Einstein found the smooth manifold of
general relativity to be much more comfortable than the roiling chaos of the
subatomic realm of quantum theory.
The birth of quantum theory, not so coincidentally, coincided with the
discovery of a zoo of subatomic particles. In 1897, Cambridge physicist J. J.
Thomson discovered the electron, a tiny charged particle weighing in at a bit
more than 1/2,000 the mass of even the smallest atom. A decade and change
later, Ernest Rutherford, running the same lab—which featured prominently
in the Hawking biopic The Theory of Everything—gave the name of ‘proton’
the much heavier charged particle he had discovered. Soon after, again in the
same laboratory, James Chadwick discovered an uncharged particle almost
as massive as the proton, the neutron. These were the three particles that,
stuck together in various configurations, made up the atoms and molecules
and all the everyday matter on Earth.
But the laws of physics had a surprise in store. In the late 1920s, Paul
Dirac, also at Cambridge, came up with an elegant equation that fixed some
of the difficulties that physicists were having with the electromagnetic force
and how it behaved when combined with the principles of quantum theory.
The equation worked beautifully, but it showed that the electron had a
doppelgänger: a particle that weighed exactly the same amount, but carried
an equal and opposite charge. (The particle was spotted shortly thereafter,
and Dirac was rewarded with the Lucasian Professorship at Cambridge—the
very one that Hawking would hold half a century later.)
Now there were more particles than chemists and physicists wanted,
more than they needed to explain the composition of matter. In 1936 came the
muon, a particle that behaved like the electron, but was quite a bit heavier
and tended to decay after a short time. “Who ever ordered that?” quipped
theoretical physicist Isidor Rabi. And the particles kept coming. Medium-
weight particles like the pion and the kaon. Heavyweight ones like the
lambda and the xi. The zoo of particles was expanding at an enormous rate. It
was more than scientists could keep up with, or make sense of. At his Nobel
lecture in 1955, physicist Willis Lamb joked, “I have heard it said that the
finder of a new elementary particle used to be rewarded with a Nobel Prize,
but such a discovery now ought to be punished with a $10,000 fine.”6
It took until the early 1960s—at the same time that Hawking was
beginning his studies at Cambridge—for scientists to begin to figure out how
the myriad particles were related. There was a hidden structure underneath
the chaos, an underlying order that, by the mid-1970s, had developed into
what is now known as the Standard Model.i
The Standard Model is a mathematical framework that describes all the
matter in the universe and how it interacts.ii This framework explains the vast
and ever-expanding zoo of subatomic beasties by positing that there are a
small handful of fundamental particles with very predictable behaviors.
On one side of the Standard Model, there are particles, the fundamental
fermions, that describe the raw materials that matter is made of: the
relatively massive quarks, which make up the protons and neutrons at the
center of every atom; the lightweight electron, which tends to live in an
atom’s periphery; as well as other, more exotic species of particle, such as
the neutrino, so light that its mass is almost undetectable. The other side of
the Standard Model describes how particles interact with each other. We
know, as just one example, that a negatively charged electron will repel a
second negatively charged electron thanks to the electromagnetic force. In the
Standard Model, this repulsion force is carried by yet another particle, a
force-carrying one known as a gauge boson. This is a particle that interacts
with both electrons and, essentially, causes them to feel each other’s
presence. Just as the Standard Model has a gauge boson that carries (or
mediates) the electromagnetic force, it also has a gauge boson that mediates
the strong force that glues the protons and neutrons together at the center of
the atom. Bosons are also responsible for a third force, known as the weak
force, that, on certain occasions, causes particles to interact by changing each
other’s identities. The Standard Model has not one, but three gauge bosons
that carry the weak force.iii
In essence, the Standard Model explains everything about matter—what
it’s made of, how it behaves, what forces it feels, and how it feels them—
with a mathematical framework that describes two kinds of particles, the
fundamental fermions and the gauge bosons. The fundamental fermions, in
various combinations and permutations, explained the nature of every single
subatomic particle that we humans have yet observed. But even more
astounding, the Standard Model describes every possible way those particles
can interact with each other, every conceivable way that they can feel each
other’s presence via three fundamental forces: the strong force, the weak
force, and the electromagnetic force. Building this framework was a
tremendous mathematical achievement, the crowning glory of late twentieth-
century physics.
But there is a problem. A big one.
There aren’t just three fundamental forces. There are four. In addition to
the strong force, the weak force, and the electromagnetic force, there is—of
course—gravity. And the Standard Model doesn’t explain gravity at all.
Particles in the Standard Model don’t feel each other’s masses; there’s no
gauge boson that explains how one particle might be gravitationally attracted
to another. With no gravity in the mathematical framework, there’s no way the
Standard Model can be considered complete. It is not a theory of everything,
just a theory of most things under many conditions.
But it’s not easy to add gravity to the Standard Model. It’s built upon a
mathematical framework that works just fine with massless particles, but
real-world particles that feel gravity and have mass tended to break the
mathematical underpinnings of the model.
This is where the Higgs boson comes in. In the early 1960s, several
scientists, including the Scottish physicist Peter Higgs, figured out how to
bridge the gap between the massless particles the mathematical framework
produced and the massive particles physicists observed in real life. This is
what is now known as the Higgs field.
Perhaps the best way to visualize the Higgs field is by analogy: imagine
that the universe is filled with some sort of viscous fluid, like honey. Some
particles, by their nature, might be blunt-shaped, clumsy things; when pushed,
they have a hard time moving through the honey. These particles resist motion
—they are massive—because they feel the effects of the honey very strongly.
Other particles might be more streamlined and would zip through the honey
with very little trouble. These particles have less mass, or may even be
massless; they don’t feel the effects of the honey very much at all. In a
vacuum, all particles, no matter their shape, would move about in the same
way; there would be no resistance. But once you add the honey, that
equivalence is broken; some particles interact more strongly with the honey
than others, and thus move with more difficulty. In the same way, a particle in
some sense starts off massless, but it acquires mass depending on how
strongly it interacts with the Higgs field.iv
As a theoretical construct, the Higgs field worked beautifully. It didn’t
truly explain where mass came from; nor did it fix the disconnect between
Einstein’s smooth picture of gravity and the lumpy, discontinuous universe of
quantum theory and interacting particles. Gravity still wasn’t a part of the
theory. But the Higgs field was an intermediate step that allowed physicists
to build a mathematical model including real-world particles with mass—
and three of the four forces affecting them. It wasn’t a complete solution, but
the Higgs mechanism was an important part of what became the Standard
Model.
However, once you assume there’s a Higgs field, there has to be a particle
associated with it. So if you believed in the plain vanilla Standard Model,
you were forced to accept that there was a so-called Higgs boson out there
waiting to be discovered.v Fail to find it, and the Standard Model would be
in trouble. And so the hunt for the Higgs began.
Physicists find exotic particles by pouring enormous amounts of energy
into a very small space: by colliding particles together at very high speeds
and looking through the spray of debris to see if they can find something new.
The higher the energy of the particle collider, the more exotic the objects that
can be found. Since nobody knew precisely how heavy or how rare the Higgs
boson was, particle physicists had only a rough sense of how powerful their
colliders had to be to spot a Higgs. Starting in the mid-1980s, they started
seeing hints—ghosts—in their machines that seemed to indicate the existence
of the Higgs. And like ghosts, they tended to be products of overactive
imaginations. After building new collider after new collider over the course
of several decades, theorists were becoming increasingly worried as the
Higgs failed to show up.
On July 4, 2012, nearly half a century after the original predictions,
scientists at CERN near Geneva announced that the search for the Higgs was
finally over.
“Physicists Find Elusive Particle Seen as Key to Universe,” blared the
New York Times on its front page. A more understated Washington Post
declared that “Scientists’ Search for Higgs Boson Yields New Subatomic
Particle.” “Scientists Might Have Found ‘God Particle,’” exclaimed the
Daily Telegraph in the United Kingdom. And, next to that, “Higgs Boson:
Prof Stephen Hawking Loses $100 Bet.”
Hawking himself had nothing to do with the discovery of the Higgs boson,
or the nitty-gritty of the Standard Model that had led to the prediction of the
Higgs. Yet on the day of the announcement, the BBC turned to Hawking to
comment on the significance of the discovery. “If the decay and other
interactions of the particle are as we expect, it will be strong evidence for
the so-called Standard Model of Particle Physics, the theory that explains all
our experiments so far,” Hawking told the reporter. “This is an important
result that should earn Peter Higgs the Nobel Prize, but it is a pity in a way
because the greatest advances in physics have come from experiments that
gave results we didn’t expect.” He then added: “For this reason, I had a bet
with Gordon Kane of [the University of Michigan] that the Higgs particle
wouldn’t be found. It seems I have just lost $100.”7
The story of the Higgs discovery, as important as it was, was an abstract
and difficult story to tell. The story of Hawking’s long-running bet gave
journalists a way to put a human face on the narrative, to tell a story that
readers would find compelling rather than confusing. Peter Higgs himself is a
shy and retiring character, modest almost to a fault—and not very good copy.
Hawking, on the other hand, was already a fixture of public life, a staple of
the science pages, and a known draw. And it probably tickled Hawking to
know that he was stealing a bit of the limelight away from Higgs.
“I didn’t really quite appreciate how much Stephen enjoyed the limelight,”
says Peter Guzzardi, Hawking’s editor for A Brief History of Time, the book
that made Hawking a household name. “I mean, he just thrived in it.” By the
time Hawking was explaining the significance of the Higgs boson to sixty
thousand people at the 2012 Paralympic Games, he had been an A-list
celebrity for a quarter century. Most of the time, he relished the attention. But
even when he did not, he couldn’t escape the public glare. “I can’t disguise
myself with a wig and dark glasses,” he wrote. “The wheelchair gives me
away.” And so his every dalliance was always in public, and this made his
personal life fodder for the tabloids. And, generally speaking, tabloids aren’t
terribly interested in the arcana of particle physics. They preferred to publish
Hawking’s quirks. And his peccadillos.8
“Renowned Physicist Stephen Hawking Frequents Sex Clubs,” trumpeted
one tabloid in February 2012:
I had reason to fear that the effort involved in sexual activity might kill
Stephen in my arms.…
The functions I fulfilled for him were all maternal rather than
marital: I fed him, I washed him, I bathed him, I dressed him, I
brushed his hair and cleaned his teeth.… It was becoming very
difficult—unnatural, even—to feel desire for someone with the body
of a Holocaust victim and the undeniable needs of an infant.13
When Jane released the second version of her memoir several years later,
this passage was softened significantly, but the essential message was the
same: she found the relentless sacrifices needed to fulfill Stephen’s needs to
be fundamentally incompatible with spousal affection.
With his divorces, Stephen had lost the most reliable mechanism for him
to get his basic needs fulfilled. Hawking had to rely upon his friends, family,
and confidants to look out for him. Otherwise, he had to impose his will on
others through the sheer force of his personality—or with his money.
Conversely, others tried to impose their will on him. The Hawking name
was extremely valuable, and a number of people close to Hawking have
described sometimes bitter conflicts behind the scenes over the control of
how he spent his wealth, his attention, and his time in the last few years of
his life. Most of these conflicts did not erupt publicly; there was one notable
exception, but even then, the details are scant.
This exception was the case of Patricia Dowdy, one of Hawking’s nurses.
Dowdy had worked for Hawking in the early 2000s, left, and then came back
several years later. In 2016, at the behest of a number of Hawking family
members, the UK Nursing and Midwifery Council opened an investigation
into Dowdy, immediately suspending her pending an inquiry. Of the twenty-
five professional standards in the council’s code of conduct, Dowdy stood
accused of violating nineteen of them. The investigation lasted for two years
and wrapped up with a hearing shortly after Hawking’s death; Dowdy
effectively lost her nursing licensevii in 2018 for “multiple misconduct
charges” related to her care of Hawking, including “financial misconduct;
dishonesty; not providing appropriate care.”14
The circumstances were irregular and rather bizarre: Though such
hearings are typically open to the public, the council made an exception in
this case and held the hearing in private. To this day, the council refuses to
release documentation regarding the proceedings, making it all but
impossible to understand the nature of the charges against Dowdy, much less
to try to get a sense of her guilt or innocence. And though Hawking’s own
papers almost certainly reveal the circumstances of the conflict that led to the
end of Dowdy’s nursing career, those documents are in the hands of the
Hawking family who accused her—and they’re not revealing anything either.
The rougher edges of Hawking’s personal life can only dimly be seen
poking out from underneath the surface, given how much effort has been
expended to keep them under wraps.
Despite the flourishing market for details about Hawking’s personal life, he
himself didn’t publish a memoir or autobiography until 2013, when he was
seventy-one years old. And when My Brief History came out, it was almost
the diametric opposite of his ex-wife’s tell-all. Reviewers were struck by its
brevity—and the author’s reticence, his “air of such studied incuriosity,
particularly about the people in his life, that it feels inveterate.”15
“As for his private life, his two marriages—swirling around them rumors
of infidelity by Hawking and his wives and of physical abuse of the disabled
scientist—and his globe-trotting experiences as a celebrity, the understated
Hawking reveals very little,” one reviewer wrote. “He admits that his failing
health and celebrity status didn’t help his relationships, but offers few details
or insights.”16
In Nature, the peer-reviewed journal where Hawking had published his
famous suggestion that black holes should eventually explode, the review
was even more brutal: “My Brief History does not do what we expect of a
memoir. It does not take the reader behind any scenes. Hawking narrates his
life non-introspectively, celebrating its triumphs and burnishing its sensitive
moments. It is a concise, gleaming portrait, not unlike those issued by the
public relations department of an institution.”17
Without any real introspection, My Brief History wasn’t likely to become
any sort of critical—or, more importantly, financial—success.viii Here, too,
Hawking’s inability to communicate at more than three words a minute
played a role in limiting the amount of new material he could produce. Much
of the book was recycled. His account of his childhood and college years
was a lightly edited version of a speech he’d given in 1987, bulked up with
additional material from a few years later. His account of the writing of A
Brief History of Time had appeared in The Independent and Popular
Science in late 1988 and early 1989. Other chapters on his scientific work
were collages of previous lectures and writings. The all-too-brief
introspective parts that reviewers (and readers) craved were the only really
novel elements in the book. Yet producing novelty—writing something new
from scratch, rather than cutting-and-pasting from prior utterances—was
what gave Hawking the most difficulty.18
Even if Hawking had found it easier to write new chapters for his
memoir, he may well not have been so inclined. He consistently tried to
maintain a dignified, humorous, and self-deprecating tone in his popular
writings. It wouldn’t have been compatible with his image to try to boost
sales by writing about the more intimate parts of his life, especially anything
that might come across as controversial or scandalous. So it shouldn’t have
come as a surprise that My Brief History gave no insights into Hawking’s
love life, or even into his politics or his religious beliefs.
Hawking made no secret of his atheism. Though he frequently referred to
God in his writings, typically as a metaphor for the underlying order of the
universe, he made it quite clear on a number of occasions that he did not
believe in a supernatural being or spiritual force.ix Indeed, the version of the
cosmological creation that he came to believe in ruled out the existence of a
creator. Yet that didn’t seem to satisfy the public’s thirst for spiritual wisdom
from the physicist; he was constantly asked about his religious views. The
total absence of God in his memoir was striking.
There were also relatively few clues about his politics. Since his youth,
Hawking had identified as a socialist—and he clearly leaned left. However,
he tended not to weigh in publicly on political matters, except on occasions
when there was a scientific angle, as with climate change or nuclear
disarmament. And in those cases, he would use his scientific credentials to
bolster his authority.
It was largely irrelevant that Hawking’s scientific expertise often had
little to do with what he opined about. His years of studying general
relativity and particle physics didn’t give him any special insight into the
nature of artificial intelligence, the need for space travel, or the desirability
of contacting alien species, for example. Yet he was more than happy to
speak at length on these subjects, as if he were representing scientific
consensus. This sometimes irked even his longtime friends, such as
astronomer Martin Rees. “[A] downside of [Hawking’s] iconic status was
that his comments attracted exaggerated attention even on topics where he
had no special expertise—for instance philosophy, or the dangers from aliens
or from intelligent machines,” he wrote in 2018. “And he was sometimes
involved in media events where his ‘script’ was written by the promoters of
causes about which he may have been ambivalent.”19
This isn’t to say that Hawking only spoke out about science-adjacent
topics. He would opine about social services for the disabled, and toward
the end of his life, he grew somewhat more outspoken about politics. He
made comments in reaction to the rise of populism in Europe and the United
States, and even backed a few Labour candidates for Parliament. He opposed
Brexit, just as a few years before that, he opposed Scottish independence
from the United Kingdom. Even so, with a few exceptions, even in purely
political matters, he tended to couch his statements in scientific and
technological terms. In 2016, he said that Brexit “would damage scientific
research in Britain”; populism was, in part, “another unintended consequence
of the global spread of the internet and social media”; and, at the same
moment that he labeled Donald Trump a demagogue, he urged the new
president to choose a different head of the Environmental Protection
Agency.20
Even when he got into a scrap with the Tories over cutting funds for the
National Health Service, Hawking took then minister of health Jeremy Hunt
to task in an op-ed as if he were a fellow physicist.x “Hunt had cherry-picked
research to justify his argument,” Hawking wrote. “For a scientist, cherry-
picking evidence is unacceptable. When public figures abuse scientific
argument, citing some studies but suppressing others to justify policies they
want to implement for other reasons, it debases scientific culture. One
consequence of this sort of behaviour is that it leads ordinary people to not
trust science at a time when scientific research and progress are more
important than ever.” Hawking rarely deviated from this pattern of
argumentation from scientific authority, and when he did, it was typically big
news.21
In mid-2013, Hawking was scheduled to appear at a conference hosted by
the then president of Israel, Shimon Peres. This got the attention of the British
Committee for the Universities of Palestine (BRICUP), an organization of
academics in the United Kingdom devoted to fighting the occupation of
Palestinian lands by promoting a boycott of Israel and through other means.
Jonathan Rosenhead, the chair of BRICUP, managed to get two dozen other
academics to write Hawking a letter declaring that they were “surprised and
deeply disappointed” that he was helping Israel to “camouflag[e] its
oppressive acts behind a cultured veneer.”22
Hawking promptly withdrew from the conference, couching it as his
“independent decision to respect the boycott” of Israel. Hawking’s employer,
Cambridge University, was apparently caught flatfooted; its spokespeople
promptly denied that their universally beloved professor had deliberately
waded into one of the most divisive political topics of the modern era.
Accusing BRICUP of “misunderstanding” the reasons for Hawking’s
withdrawal, one university spokesperson insisted that “Professor Hawking
will not be attending the conference in Israel in June for health reasons—his
doctors have advised against him flying.” But it was Cambridge that had
misunderstood; Hawking canceled because of the boycott.23
Hawking’s controversial stand meant that he would receive some
extremely negative publicity—something he had mostly avoided throughout
his life. While Palestinians and boycott supporters praised the physicist,
critics attacked him. Several leveled accusations of hypocrisy against
Hawking because his speech synthesizer relied on an Israeli-made computer
chip. One furious professor at the University of Haifa even suggested that
Hawking should get “a free trip on the Achille Lauro,” the cruise ship from
which hijackers had infamously thrown a passenger in a wheelchair
overboard.24
This was an extreme reaction to a rare exception. Hawking didn’t shy
away from controversy, but when he got into a scrap, it was almost always
about something scientific—so it was generally only fellow scientists who
got really annoyed with him. Scientists like Peter Higgs.
In 2012, when the Higgs boson was found at the huge Large Hadron Collider
(LHC) at CERN, newspapers happily reported that Hawking had lost his
$100 bet with physicist Gordy Kane. What they didn’t report was that it was
his third bet with Kane—and Hawking had won the other two.
Back in the mid-1990s, the tunnels at CERN were home to a different kind
of particle accelerator known as the Large Electron-Positron Collider (LEP).
The LEP was weaker than the LHC, but at the time, physicists were
predicting that it would find the Higgs. Hawking was skeptical; his own
calculations seemed to imply that on the very smallest scales—when you
zoom in far enough to see the frothy, discontinuous substrate where the Higgs
particle belongs—tiny black holes were constantly popping in and out of
existence. This may seem like a bizarre idea, but the weirdness of the
subatomic world means that such things are eminently possible. These tiny
black holes, Hawking argued, would mask the existence of the Higgs boson
—they would make it impossible to observe any trace of a Higgs. Most
scientists didn’t buy Hawking’s argument, although fellow physicist Malcolm
MacCallum described it at the time as “wild and provocative.”25
It provoked Gordy Kane, at the very least. A theoretical particle physicist
at the University of Michigan, Kane was extremely sanguine about the
prospects at the LEP. He thought not only that it would likely spot the Higgs,
but that it would discover hints of new exotic particles whose existence lie
beyond the predictions of the Standard Model. So he bet Hawking $100 that
the LEP would find the Higgs.
By the year 2000, things were looking grim for the Higgs search. After
roughly a decade of operation, the accelerator had found scant evidence of
anything resembling a Higgs, and the LEP was scheduled to be demolished to
make way for the more powerful LHC. At the very last minute, just as the
LEP was about to be shut down, scientists claimed to see a blip that might be
a Higgs-like particle.xi The LEP’s execution was stayed a month to see if
scientists could gather more evidence for a discovery, but no dice. The Higgs
had not been found—and theoretical particle physicists were beside
themselves with frustration. Hawking proudly claimed his $100 from Kane.
The ordinarily retiring Peter Higgs snapped, rubbishing Hawking’s theory
that the boson would never be found. “It is very difficult to engage
[Hawking] in discussion, and so he has got away with pronouncements in a
way that other people would not,” Higgs told The Scotsman. “His celebrity
status gives him instant credibility that others do not have.”26
“I am surprised by the depth of feeling in Higgs’ remarks. I would hope
one could discuss scientific issues without personal remarks,” Hawking
reportedly responded, before turning the knife in the wound. “Higgs has got it
wrong. I did not bet that the Higgs Boson doesn’t exist, just that it would not
be discovered at LEP and I have already won the bet.”27
Just as the LEP was shutting down, the Tevatron, a newly upgraded
accelerator at Fermilab in the United States was starting up. Once again,
Hawking and Kane took opposing sides of the wager: Kane, firmly believing
in the Higgs, bet that the Tevatron would find it; Hawking, still doubting
whether the Higgs would ever be spotted, bet against.
Running a particle accelerator is an extremely expensive affair, even once
the equipment’s already built and in place. It sucks up an enormous amount of
power, not just to get the particles moving close to the speed of light, but also
to power and cool the magnets that steer the particle beams, and to maintain a
vacuum in the miles upon miles of tubing the particles travel in. By the mid-
2000s, the US Department of Energy, which ran Fermilab, was sending out
strong signals that the Tevatron was going to be the last US collider on the
frontier of particle physics. If the Tevatron didn’t find the Higgs, that was it
as far as America was concerned—no bigger, better collider was going to be
built. And as the end of the decade drew closer, it became clearer and
clearer that the Higgs wasn’t going to reveal itself; there were strong hints of
where one might be hiding, but nothing that physicists could remotely call a
discovery. It was looking like the Tevatron would soon be shut down, and
Hawking would once again win his bet with Kane.
The situation was looking desperate. But the game wasn’t over yet. There
was one last hope to find the Higgs, and one last bet for Kane and Hawking
to make. The LEP had been torn down to make way for a more powerful
accelerator in the same tunnels—the LHC. If the LHC didn’t find the Higgs,
then it was game over, at least for the foreseeable future. It was highly
unlikely that another machine would ever be built that could succeed where
the LHC failed. So in late 2008, with the LHC finally ready to be turned on,
Higgs hunters looked on with anticipation—and trepidation.
“I think it will be much more exciting if we don’t find the Higgs. That will
show something is wrong, and we need to think again,” Hawking told the
BBC the day before the accelerator started collecting data. Because
Hawking’s synthesized voice doesn’t inflect to show emotion, it was hard to
tell whether Hawking was being impish when he added, “I have a bet of
$100 that we won’t find the Higgs.” Given the heightened emotions in the
Higgs community at the time, he almost certainly knew he would provoke a
reaction.28
The next day, at a press conference celebrating the startup of the LHC,
Peter Higgs lashed out, deriding Hawking’s theory that the Higgs boson
would never be found. “I have to confess that I haven’t read the paper in
which Stephen Hawking makes this claim. But I have read one he wrote,
which I think is the basis for the kind of calculation he does. And frankly, I
don’t think the way he does it is good enough,” Higgs grumbled. “My
understanding is that he puts together theories in particle physics with
gravity… in a way which no theoretical particle physicist would believe is
the correct theory.… I am very doubtful about his calculations.” Apparently,
other scientists on the panel then “moved swiftly to cut off the discussion”
before things got out of hand.29
It would be four more years before the LHC finally produced enough
evidence of the Higgs for Hawking to concede his final bet with Gordy Kane.
But in the interim, Hawking seemed to enjoy tweaking Higgs and those at the
LHC who were searching for the Higgs boson. On several occasions,
Hawking half-heartedly joked that the accelerator could produce miniature
black holes that would earn not Higgs, but Hawking, a Nobel Prize.xii And
when he finally conceded that he was wrong in 2012, admitting that Peter
Higgs deserved a Nobel, he immediately followed up with his statement that
“it was a pity in a way” that the Higgs boson had been found.
Peter Higgs would never say so, but on this point at least, Hawking was
right. Finding the Higgs boson marked the completion of the Standard Model
of particle physics; it was the last missing piece in a very successful theory.
But the theory is inadequate. Even though the Higgs allows particles to have
mass, it doesn’t explain gravity. For that, scientists need to build a bigger,
better model, something that goes beyond the Standard Model to describe not
just the strong, weak, and electromagnetic forces, but the gravitational force
as well. Physicists need to start seeing evidence of phenomena that can’t be
explained by the existing Standard Model to figure out where to extend it and
improve it. It is precisely where the old models break down that the new
physics begins. If the old models don’t fail, there’s nowhere to start.
It’s for this reason Hawking had spent his life seeking out boundaries
where theories clash and break down—at the edge of a black hole, at the
birth of the universe, because it is those boundaries that give scientists
fleeting clues to truths even more profound than the ones we already know.
By pouring enormous energy into a very small space, the LHC had the
potential to provide just such a boundary. But it failed to do so, even as it
succeeded in producing the Higgs. The accelerator didn’t produce any hints
of beyond-Standard-Model physics, no clash of theories that begets new
knowledge. So, to Hawking, who had spent much of his life hoping for a
“theory of everything” that would reconcile the frothy discontinuous
subatomic realm and the smooth one of general relativity, the LHC was a
disappointment of a lifetime.
Footnotes
i For more details about the Standard Model and its development, see my
book Alpha and Omega, chaps. 8 and 9.
ii There is reason to believe that there are kinds of matter that aren’t
accounted for by the Standard Model, but we haven’t discovered them yet.
See my Alpha and Omega, chaps. 7 and 10.
iii One of the main triumphs of the Standard Model is to show that the
electromagnetic force and the weak force are really, fundamentally, the same
thing, even though they look very different. It’s somewhat akin to how ice and
liquid water seem very different until you heat them up to a high enough
temperature. Similarly, the electromagnetic force and weak force become
indistinguishable when temperatures are very, very high—as they were soon
after the Big Bang.
iv This is obviously an oversimplification, and it’s somewhat unsatisfying for
a number of reasons, not least of which is that most of the mass of ordinary
matter comes from a different mechanism. It’s also somewhat misleading:
there are four Higgs fields (and associated bosons), and we can only observe
one of them, for reasons that have to do with the same theory that ties together
the electromagnetic and weak forces.
v It’s a different variety of boson from the ones that carry the
electromagnetic, weak, and strong forces, but has some properties in
common.
vi Despite his disability, Hawking’s sex drive was essentially intact.
vii Technically, she was “struck off” the nursing registry.
viii Hawking’s financial situation improved temporarily in 2013 when he
won a $3 million prize sponsored by Russian billionaire Yuri Milner. Milner
and Hawking jointly made headlines again three years later when Milner
started funding, and Hawking publicly supported, a rather odd proposal to
send tiny laser-propelled spacecraft to Alpha Centauri.
ix The only significant suggestion to the contrary was made by his second
wife, Elaine. If she is to be believed, during their marriage she regularly took
Stephen to services at St. Barnabas in Cambridge, where “the prayers,
hymns, and bible readings often evoked a lot of quiet weeping from him.” In
addition, she said, “At home, he read the Bible. He usually asked for the Old
Testament stories and [we] often prayed together.” Elaine Hawking,
“Baptism Testimony of Elaine Hawking,” April 2018,
https://archive.org/details/chipping-campden-baptist-church-513231/2018-
04-01-Baptism-Testimony-Elaine-Hawking-Elaine-Hawking-46453311.mp3.
x Toward the end of his life, Hawking’s worries that the Conservative Party
was destroying the NHS drove him to become much more publicly partisan,
to the point of endorsing a Labour candidate for Parliament in 2017.
xi In reality, it was a budgeton, an evanescent particle whose existence is
fleetingly glimpsed by every accelerator that’s in imminent danger of being
shut down.
xii Even after he conceded the bet, Hawking seemed to enjoy popularizing a
theory that measurements of the Higgs boson might imply that the universe is
in some sense unstable.
CHAPTER 4
The time-travel stunt had Stephen Hawking written all over it; his wit and
humor combined with his unflinching confidence about the way the universe
worked would mesmerize an audience every time. Little doubt it was a
moment of authentic Hawking shining through in an otherwise humdrum
Hawking-branded Discovery Channel special.
It was getting harder and harder to find the authentic Hawking underneath
the persona, to distinguish the scientist’s original thoughts and words from
those that were thought and written for him by cadres of script writers and
coauthors. Even in his science, it was getting increasingly difficult to figure
out where he ended and his collaborators began. Stephen Hawking was as
much a brand as he was a person.
In this, Hawking was little different from many celebrities, even from
many scientists, who receive (sometimes disproportionate) credit for the
work performed by the assistants in their lab or the graduate students under
their tutelage. Hawking’s situation was more extreme because of his
condition; almost every function of his life—eating and defecating, much less
writing and preparing lectures—was a collaborative effort. Hawking had the
ability, and frequently the necessity, to outsource some of the basic functions
of the self. It could be hard to discern the line between Hawking the person
and Hawking the collective creation. The brand was so famous that it was
sometimes hard to remember there was a human, and an extremely ill one,
underneath it all.
There were two things almost always true about a public Stephen Hawking
lecture. First, it was very well attended—every seat taken, and if the fire
marshals weren’t a force to be reckoned with, large clots of people near the
exits and in the aisles, craning their necks to get their first view of the
celebrity physicist. The second was the hush that fell over the crowd as he
wheeled out, accompanied by—or later in life, steered by—an attendant.
There was typically a rustle of tension in the crowd as the attendant fiddled
with the equipment and Hawking prepared to speak. And then came the first
words, intoned exactly the same way each time, as if in a ritual: “Can you
hear me?” The tension broke as the audience applauded, and the physicist
triggered the voicebox to begin pronouncing the first words of his speech.
“Sometimes there was thirty, forty seconds of pure silence,” says
Christophe Galfard, a PhD student of Hawking’s who, inspired by his
adviser, become a science popularizer after he graduated. “For me, it was the
silence within those forty seconds that made it so… that’s what triggered my
wish to pursue that road.”3
Like many speakers on the public lecture circuit, Hawking had a set of
stock talks that he would draw from, and he would occasionally tweak them
or remix them, updating the science when appropriate, or adding new
imagery. His talks were filled with fragments of text from his earlier books,
and his later books contained lots of text from his talks. In fact, because all of
Hawking’s communications were mediated by his computer—the drafts of
his books, the speeches he gave and the ones he never did, the answers to
questioning by a colleague or an audience member—his utterances were
never evanescent, but semipermanent, stored in digital memory. Every time
he spoke or wrote, those phrases survived for decades, mutating, merging,
and producing offspring every time Hawking searched his memory banks for
something to say.
Hélène Mialet, a sociologist of science, wrote of the travails this situation
caused for a librarian’s assistant going through the physicist’s papers.
Typically, an archivist assesses the value of a document by its uniqueness:
Hawking was not the only public lecturer reusing his talks and papers—far
from it. But the consequences of his illness were that any information in his
brain that he wanted to transmit to the outside world had to flow through a
digital channel—a channel that obscured the original thought that created that
information in the first place. As Mialet put it:
For years now, he has had to do everything through the computer. This
is true not only for Hawking, but increasingly for all scholars.
Hawking’s particular condition… makes visible the practices of
intellectuals who cut and paste from different documents to compose
new talks, who recycle the same talk for different audiences, who
transform their talks into articles or publish talks, interviews, and
conferences as books. In other words, the drafting process that would
normally take us closer to the “origin” of the process of creation—that
is, to “the author”—lead[s] instead to a hall of mirrors.5
Interestingly, the questions are always more or less the same.… “Well,
basically,” the assistant explains, “what they are saying is: ‘Do you
believe that we’re going to find a [theory of everything]?’ And
Stephen has given a public lecture, a number of times… called ‘The
Millennium Lecture,’ which he gave to Bill Clinton, where he is
saying in twenty-five years. What Stephen will probably do to answer
that question is just take the last paragraph of his public lecture and
say it.”6
The title of The Grand Design evoked a hope that physicists had been
harboring for more than half a century: sometime soon, scientists would
figure out a simple set of equations; a beautiful, self-contained list of rules
that governed how all matter and energy in the universe behave under all
circumstances and on every scale. All the different kinds of particles, all the
forces in the cosmos, all these would be described completely by these
equations. This is the mythical “theory of everything”—the white whale of
theoretical physicists since Einstein. This set of rules would essentially be
the blueprint for the universe. It would be the Grand Design—the constraints
that even the creator had to obey.
Physicist Richard Feynman once likened the search for the rules of the
universe to trying to figure out the rules of a game of chess by watching it
being played—but you’re only able to see a few squares of the board rather
than the whole thing at once. Given a short time, you might figure out that a
pawn only moves forward, that each side has a bishop that stays only on the
white squares and another that stays on the black ones, and a few basic rules
of the game. But then, “you suddenly discover one day in some chess game
that the bishop doesn’t maintain its color, it changes its color. Only later do
you discover the new possibility that the bishop is captured and that a pawn
went all the way down to the queen’s end to produce a new bishop. That
could happen, but you didn’t know it.” Feynman explained. “And so it’s very
analogous to the way our laws are. They sometimes look positive, they keep
on working, and all of a sudden, some little gimmick shows that they’re
wrong—and then we have to investigate the conditions under which this
bishop changed color.”14
Physicists posit a set of rules about how the world works, and then they
make observations, over and over again, to see whether those rules are
holding as expected. Most of the time, they do—but when they don’t, there’s
the possibility of getting a deeper, better understanding than before.
For example, by the end of the eighteenth century, scientists had divined a
set of rules that seemed to be working pretty well. Atomic theory described
the composition of matter: it was made of invisible, uncuttable atoms that had
various properties. Newton’s law of gravitation described how matter was
affected by (and affected) the gravitational force, and Maxwell’s equations
described electricity and magnetism. In the early twentieth century, though,
scientists were seeing that the atom itself wasn’t uncuttable, but composed of
pieces of different sorts (electrons, neutrons, and protons), and that a
different sort of force altogether—the weak force—was involved in
radioactive decays. By the 1970s, the rulebook had gotten yet more intricate;
protons and neutrons were themselves composed of smaller particles,
quarks, and were bound together with another kind of force, now known as
the strong force.
It might seem that the rules of the game have been getting more and more
byzantine at each iteration. But from a physicist’s point of view, it’s just the
opposite: the rules are getting simpler, thanks to unification. Scientists
realized that the electric force and magnetic force are really different aspects
of the same thing; the weak force, too, at high enough energies, is also
governed by the same fundamental laws. Similarly, the whole particle zoo,
which seemingly produces an endless variety of different subatomic
creatures, can be explained with a small handful of fundamental particles. As
Feynman put it, “in the case of the chess game, the rules become more
complicated as you go along, but in the physics when you discover new
things, it becomes more simple.” Perhaps it’s a simplicity that’s only truly
appreciated by theoretical physicists, but the mathematical framework of the
Standard Model is remarkably simple—with just a few assumptions, it’s
possible to describe all the kinds of matter that scientists have ever
encountered, and how they are affected by three of the four fundamental
forces: electromagnetism, the weak force, and the strong force. However,
since gravity isn’t included, at best, you can have a theory of almost
everything.15
So physicists watch and wait, observing the universe through their
particle colliders and their telescopes, hoping to see Nature breaking the
rules so they can come up with newer, deeper, and hopefully simpler sets of
rules that explain more. Perhaps, even, a theory of everything, a set of rules
that would explain the behavior of all matter and energy in the universe. But
such a theory would require a new mathematical framework—and new
assumptions about the physical world that go along with that framework.
What, precisely could that framework be?
By the time Hawking wrote The Grand Design, he had come to believe
that a mathematical construct known as M-theory was the only real contender
for that framework. M-theory wasn’t really a theory—yet. It was a
hypothetical construct, an inkling of a mathematical framework that theorists
knew existed but couldn’t yet describe very well. And to Hawking—as well
as many other physicists—M-theory was the best shot at going beyond the
Standard Model to explain all the matter and forces in the universe.
However, M-theory—as potential theories of everything are wont to do—
comes with some philosophical baggage that needs to be explained. For
example, M-theory seems to imply the potential (or worse, the actual)
existence of multiple different universes with different physical laws.
Hawking spent a good portion of the book tackling this implication and
explaining why we inhabit the universe that we do among all the myriad
possibilities.vii
Unfortunately, Hawking was fairly late to jump on the M-theory
bandwagon, and there had already been several popular books written on the
subject. So those reviewers who could look beyond the “Hawking denies
God” storyline tended to be disappointed. “Even allowing for the need for
the pleasures of digression, there is too much padding and too much recycling
of long-stale material,” wrote Graham Farmelo, a physicist and writer of
popular math and physics books. “It gives me no pleasure at all to say that I
doubt whether The Grand Design would have been published if Hawking’s
name were not on the cover.”16
Nevertheless, Hawking’s name was on the cover. Upon its release, The
Grand Design immediately shot up to number one on the New York Times
Best Seller list. Its sales were almost certainly helped by controversy. It
wasn’t just Hawking’s stance on God that got critics clicking their tongues,
but also his swipe at one of his perennial targets. “Philosophy is dead,” said
The Grand Design on its opening page. “Philosophy has not kept up with
modern developments in science, particularly physics.” It was a statement
that seemed calculated to irk a certain segment of academia.
Though successful by any reasonable measure, The Grand Design was no
Brief History of Time. For most authors, eight weeks on the Times Best
Seller list would be the achievement of a lifetime. But Brief History had
been there for more than two years solid. Grand Design just didn’t come
together in the same way. Even Hawking’s distinctive sense of humor—wry,
sometimes wicked, and yet self-deprecating—came across as flat. “Borscht
belt,” complained the reviewer for the New York Times, who seemed to pin
the blame on Mlodinow rather than Hawking. Yet it was far from clear where
Mlodinow’s contribution ended and Hawking’s began.17
At nearly seventy years old, Hawking was no longer taking new graduate
students. Over his career, he had been the adviser on more than forty PhD
theses, and his last advisee graduated in 2010. But Hawking kept working
with his peers and publishing scientific papers, most of which were
coauthored with two longtime collaborators, James Hartle at the University
of California at Santa Barbara and Thomas Hertog at KU Leuven in Belgium,
and which mostly dealt with a subject that Hawking had been grappling with
ever since his own PhD thesis nearly a half century earlier. It, too, had to do
with time. Specifically, how it began.
The universe, as we now know, had a beginning. About 13.7 billion years
ago, the universe came to be in a Big Bang. There was a moment of creation
when the very fabric of space and time came into being. Since then, that
fabric has been stretching—expanding, making galaxies rush away from one
another. If you were able to travel backward in time, you’d be able to see the
expansion running in reverse: the fabric of the universe would shrink and
shrink, getting tighter and tighter until…
Until what? This is one of the questions that Hawking spent his career
trying to answer. His PhD thesis came to the conclusion that there had to be a
breakdown in the laws of general relativity at the moment of the Big Bang,
that the fabric no longer could behave like the smooth manifold the laws of
relativity require in order to work properly. However, for the last quarter
century of his life, Hawking, along with Hartle, had come up with a clever
dodge, a way to preserve the smoothness of the spacetime fabric even at the
very moment of the Big Bang. Hawking was extremely proud of this solution,
which came to be known as the Hawking-Hartle no-boundary proposal.
The no-boundary proposal was a powerful but controversial idea, and
since the early 1980s, Hawking firmly believed that it explained what had
happened to spacetime at the beginning of the universe. It also became the
scientific foundation from which he felt he could dispense with the idea of a
creator for the universe.
Footnotes
i One possibility that didn’t occur to Hawking was that the lack of attendance
might be because the champagne was plonk.
ii That is, the distance, d, between two objects in Newtonian space is
CONCESSIONS (2004–2007)
Even as his personal life was in turmoil—his ex-wife and children all but
estranged, police looking into allegations of abuse at home, frequent visits to
the hospital as his health waxed and waned—Hawking found solace at work,
at the Department of Applied Mathematics and Theoretical Physics at
Cambridge University, which he visited almost every day. His office,
decorated with his awards and with a large poster of Marilyn Monroe,
offered escape. As he liked to say, “Although I cannot move and I have to
speak from a computer, in my mind, I am free.”13
Hawking bristled at being considered vulnerable—unable to move his
muscles, he shaped the world around him through sheer force of will. And
that will was formidable, if sometimes hard to live with. “If you are a
disabled person and you are going to do something with your life, then there
is a measure of egocentricity. If you are reliant on other people all the time,
you’ve got to have that force of ego that will persuade people to do things for
you,” his daughter, Lucy, recalled. “But then, if I were in that situation, I
doubt whether I’d have the determination to do as much as my dad does and
to keep going. He is incredibly tenacious; a very stubborn person. Once he
sets his mind on something, he’s like an ocean liner: he doesn’t change
course.”14
But change course he could—and did, even on fundamental matters. For
much of his career, Hawking had plowed on, full steam ahead, with the
conviction that there was a fundamental problem with one of the two great
physical theories of the twentieth century, quantum theory. Until he convinced
himself that he was wrong.
Quantum theory, like relativity, was born at the turn of the century, and its
origins had to do with equations that weren’t working as they should. And
once again, the equations had to do with light.
In the 1800s, physicists were firmly convinced that light was a wave.
That is, light behaved like smooth, undulating ripples on the surface of a
pond rather than like a collection of tiny, discrete objects—corpuscles or
particles. And the equations that dictate how light moves have exactly the
same form as equations that describe the motion of waves in the ocean.
However, by the turn of the century, physicists began to realize that
treating light as a wave didn’t always work. For example, in 1900, an
eminent British physicist, Lord Rayleigh, used wave equations to calculate
how light would behave when trapped in a reflective box at a given
temperature. He found, to his chagrin, that no matter the temperature, no
matter the size of the box, the wave equations dictate that the cavity inside
contains an infinite amount of light energy. That simply couldn’t be.
Something was going wrong.
Nobody knew it yet, but this was the end of the era of classical physics; a
new and bizarre theoretical framework was slouching toward Germany to be
born.
That framework was quantum physics. In December 1900, a forty-two-
year-old German physicist, Max Planck, was attacking roughly the same
problem that Rayleigh was, but from a slightly different direction. Instead of
imagining an empty box, Planck started with a lump of matter and tried to
calculate what sort of light it could absorb and emit. The equations that
resulted gave an answer that baffled Planck at first. They seemed to imply
that the lump of material could only absorb or emit light that came in discrete
packets—packets that had a minimum size—rather than as a smooth,
continuous wave. Light wasn’t behaving like a smooth, continuous wave.
Planck soon realized that with this assumption about light, this “quantum
hypothesis,” his calculations suddenly matched what scientists were seeing
in the laboratory when they measured how much light was emitted by an
object at a given temperature. The equations of quantized light had succeeded
where the equations of continuous wave-like light had failed.15
Planck later described his quantum hypothesis as an “act of desperation,”
a mathematical kludge that would make the equations of physics start
working correctly again. He didn’t realize at the time that he had stumbled
upon a fundamental property of the universe on very small scales: nature
wasn’t continuous, but discrete. And in 1905, patent clerk Albert Einstein
explained a bizarre property of metals—how certain colors of light would
cause them to spit out electrons when others would not—by applying the
quantum hypothesis. But unlike Planck, Einstein suggested that the quantum
hypothesis held for all kinds of light, everywhere in the universe, not just the
sort of light that shines from a body at a given temperature. Then, in 1913,
Niels Bohr realized that the quantum hypothesis applied not just to light, but
to matter, too: by assuming that the electron in the hydrogen atom could only
take on certain energies and not others, Bohr was seemingly able to explain,
with great precision, the colors of light that hydrogen would absorb and emit.
The quantum hypothesis had quickly become the key to understanding the
behavior of objects on the very small scale; there was something
fundamentally discontinuous about electrons and atoms and packets of light.
Planck, Bohr, and Einstein all won the Nobel Prize for their quantum-
theoretical insights.i
However, it wasn’t until the mid-1920s that physicists figured out just
how weird a quantized universe really was. Planck’s desperate act had
loosed upon the world a totally different conception of the fabric of reality.
Quantum theory is founded on our inability to predict with certainty
precisely how an object is going to behave. This inability, this uncertainty, is
woven into the very fabric of the mathematics of the theory. On short time
scales and in small regions of space, that mathematical framework paints a
picture of nature that is unpredictable and chaotic. Questions that have simple
answers in classical physics—Where is an object? How fast is it moving?
How much energy does it have?—can have fuzzy answers or even no
accessible answers at all in the quantum realm. Or, another way of looking at
it: in quantum theory, nature limits the information it does and doesn’t reveal
to an observer.
This is a very strange property, making quantum theory unlike any
scientific theory that came before it. And it has a flip side: just as the
mathematics of quantum theory treats information about the natural world as
something that can’t always be extracted, it treats that information as
inviolable. That is, this information can be transferred from place to place, it
can be split up and scattered, it can be transformed, but it can never truly be
lost. A neutron never can lose touch with its “neutronness” even if it gets
split apart or transformed (say, into a proton, an electron, and an antineutrino,
as often happens). If a scientist painstakingly collected all the parts—the
proton, electron, antineutrino, and energy—in theory, she would be able to
reconstruct the neutron from the debris it left behind when it decayed.
It’s not just a matter of different philosophies, or different mathematical
approaches. It’s a fundamental incompatibility between quantum theory’s
picture of how nature works and relativity theory’s. In relativity, the canvas
of spacetime is smooth and gently curved. In quantum theory, the canvas is
rough and foamy on small scales. Quantum theory dictates that information
about an object is never truly lost. Relativity makes no such promise. Indeed,
the so-called no-hair theorems seem to imply otherwise. This is the issue at
the heart of the black-hole information paradox.
The phrase “black holes have no hair” is physicist shorthand for the
assertion that black holes are almost entirely featureless; other than the
collapsed star’s mass, charge, and rotation rate, there is nothing that an
outside observer can determine. It is impossible to measure the black hole
and figure out what it was made of—how much of its mass came from
swallowed neutrons or swallowed protons and electrons, how those neutrons
or protons and electrons were moving before they fell into the black hole, or
anything like that. All that information is lost; once the matter falls past the
event horizon, it is gone. Not preserved, like quantum mechanics said it was,
but totally destroyed. To the people who studied general relativity—Stephen
Hawking chief among them—a black hole was a cosmic eraser that
permanently wiped out information.
This was a belief that struck at the heart of quantum theory, a belief that
Hawking had held for three decades. And yet, even as the controversy about
his home life swirled about him, the ocean liner Hawking had made a sudden
and drastic U-turn.
Perhaps even more disappointing were the sales of his brand-new book. A
Briefer History of Time, which came out in late 2005, was supposed to be a
more accessible, updated version of his 1988 best seller. And unlike the
original, Hawking had formally contracted with someone who could help
with the prose. Leonard Mlodinow, the Caltech physicist/author, wasn’t
given full coauthorship, but right underneath Hawking’s name on the cover
(in a smaller font, naturally), it stated that Hawking had written the book
“with Leonard Mlodinow.” Unfortunately for both men, the reviews were far
from stellar.
“Long story short: the only way Hawking could have dumbed it down any
further would have been to enlist Paris Hilton as a co-author, which would
have cost him a certain amount of academic cred,” snarked one wit in
Maclean’s. “But then again, it would have made the chapter on the Big Bang
far more enrapturing.” A more thoughtful review, in Nature, had a different
take: “With the briefer version, I feel the baby has been thrown out with the
bathwater,” as the book lacked the “charming incomprehensibility” of a
glimpse inside the mind of one of the world’s leading physicists, the
reviewer wrote. “It is just another run-of-the-mill popular science book on
modern physics,” with many of the topics in the book having been covered
more skillfully by other authors.23
But the review that probably stung the most was the one in The Times. “It
is interesting that the impetus for A Briefer History of Time was partly to
make it more accessible by removing the maths of the original. This would
be fine, except that Hawking is not very good at words. His language lacks
the clarity and explanatory force of masters such as Brian Greene and Paul
Davies.”24
A Briefer History of Time didn’t crack the best-seller lists. If it didn’t
sell, then his other books in the works didn’t have much of a chance.
Shortly after Briefer History came out with his usual publisher, Bantam
Books, Hawking started working on God Created the Integers with Running
Press, an imprint out of Philadelphia. The idea of the book was to give
readers a sample of key mathematical texts through the ages—such as
Euclid’s Elements, Archimedes’ Sand Reckoner, and Descartes’ Geometry.
Hawking would organize and edit the works and write short introductions to
each to put the mathematics and its author in context. It probably wasn’t the
sort of book that would sell great quantities under the best of circumstances,
but there was a chance that Hawking’s name-draw would generate some
sales and make him and his publisher a little bit of money.
However, if A Briefer History of Time wasn’t Hawking at his best, God
Created the Integers was Hawking at his worst.
The book was carelessly edited and the core texts had lots of minor
issues; the introductions had more significant errors that in some cases
seemed to be the result of patchwriting.
Patchwriting is a form of plagiarism in which an author cleaves very
closely to a text, changing some words around so that it’s not “copied” in the
ordinary sense of the word—but is still an improperly cited and badly
digested regurgitation of someone else’s work. It’s not always easy to spot—
and there’s some latitude in interpretation—but patchwriting leaves some
telltale signs: bizarre mistakes that stem from a slight misunderstanding of the
original text; odd leaps of logic or in the focus of the narrative that are
caused by omissions or distortions of the source text; phrasings that parallel
the source material for long passages without explanation; attention to details
that seem quirky or irrelevant outside of their original context. And the
introduction-and-context parts of God Created the Integers suggested that
large sections had been patchwritten from a number of different sources.
In the Descartes section, Hawking described the philosopher’s death:
For the first two years after their marriage, they lived in a house
called “College View,” which was about a ten-minute walk to the
college. It was convenient, but soon became too crowded for the
Booles.… [T]hey moved to a rented house in the village of Blackrock,
four miles from the college but only half a mile from the railway
station. The Booles loved this house. It had a wonderful view of
Cork’s magnificent harbor.27
For the first two years of their marriage, they lived, probably in rented
accommodation, in a house called “College View.”… Here Boole was
within ten minutes walking distance from the college in an area much
favored by the rich merchant classes of the city.… Early in 1857, the
Booles decided that their expanding family needed more room so they
moved to a modest rented house in Castle Road, a short distance from
the village of Blackrock, some four miles from Queen’s college. The
house stood overlooking the sea with a good view of Cork’s
magnificent harbor.… [T]he little Blackrock station was about half a
mile from the house.28
There are lots of other examples from biographies and other similar (uncited)
sources—The Dictionary of Scientific Biography seemed to be the
inspiration for more than one section. None of these are damning on their
own (or even, arguably, damning in sum), but the writing process was
inarguably below the quality that one would expect from a major author. The
errors that crept in, the carelessness of phrasing, the seeming slavishness to a
small handful of sources: all of these led to a book that not only failed to
make Hawking very much money, but risked tarnishing his reputation as a
communicator of science.
However, it is almost certain that Hawking himself did not write the
passages in question. Merely holding a book to patchwrite from was beyond
the physicist’s capability by the time he “wrote” God Created the Integers;
it’s hard to imagine Hawking composing text at three words a minute while
having an assistant help him pore over the source material. Far, far more
efficient would be to have one of his many assistants do the work. Hawking
would then put his name on it in hopes of selling more copies. “Well, the
Running Press books, he didn’t write those books,” says Al Zuckerman,
Hawking’s agent at the time. Running Press had gotten Hawking’s name on
not just God Created the Integers, but also a similar (and rather better-
executed) biographies-plus-primary-sources book about physics, On the
Shoulders of Giants, a few years prior. “Those books were compendiums,
for which he wrote the introductions, and he in fact didn’t even write those
introductions.… Hawking’s name was put on the books to make money.”
Gil King, who ghosted five of the biographical essays in On the
Shoulders of Giants for Hawking, got the impression that Hawking wrote the
introduction and little else.iv “I think Stephen Hawking wrote an original
introductory essay, but they said, ‘You know, he’s not really going to be able
to write the biographies of Galileo and Newton,” King recalls. “They needed
just like a basic introduction of who these physicists were. They had me
working with a real physicist… [and] I basically just did all the biographical
stuff. There was really no need for me to interact at all with Hawking; I don’t
really even know how much he wanted to be involved with this project. So it
was sort of like a way to get a product out there, I guess. And so that’s what I
did.” King only heard from Hawking’s people once or twice during the entire
process. “It was sort of a work-for-hire thing. I finished it fairly quickly. And
I didn’t really interact. I think the most I ever got was an email from Stephen
Hawking’s assistant. It was a basic question about, you know, deadline or
word count or something like that.” After completing his essays, King
forwarded his work on to the publisher to get Hawking’s name stamped on
it.29
In this, Hawking is little different from many modern artists—such as
Andy Warhol or Jeff Koons—who have had other people manufacture their
work. (“I’m the idea person,” Koons told the Journal of Contemporary Art.
“I’m not physically involved in the production.”) When it came to
patchwriting in God Created the Integers, it seems likely that one of
Hawking’s ghostwriters let the physicist down.30
Hawking took sole credit for God Created the Integers: his name was the
only one on the cover and in the author biography. (“Stephen Hawking is
considered the most brilliant theoretical physicist since Einstein…”) In the
acknowledgments, he does, however, extend special thanks to a number of
people, some of whom were ghostwriting for him.v But whoever was
responsible for the major problems with the book, Hawking himself cut
dangerous corners when he agreed to put his name on it.31
The Hawking name was a valuable commodity. The physicist was more
than just a person: he had become a brand, a symbol. Something to monetize.
And Stephen Hawking himself only got a small fraction of what his brand
was worth.
Not long before A Briefer History of Time came out with Bantam, a
different publisher, New Millennium, published an embarrassingly slim
volume: The Theory of Everything: The Origin and Fate of the Universe.
As with all his books, Stephen Hawking’s name was plastered on the front in
huge type—he had long been one of those rare authors whose name was
always a stronger selling point than the title of the book. And as with most of
his books, its publisher tried to extract the maximum value out of it by
remixing and rebranding it; within a year, a bulkier “Illustrated” version—
still under two hundred pages—was hitting the shelves. “Although the book
is short, it is full of information,” wrote one reviewer. “And despite
Hawking’s efforts to keep it simple, the going gets pretty dense sometimes.”32
But Hawking didn’t write the book, at least not in the ordinary sense.
Despite the huge “Stephen W. Hawking” emblazoned on the front cover,
Hawking had not written anything for New Millennium; he didn’t want them
to publish a book in his name, and they did it over his objections.vi
The Theory of Everything was old Hawking material—transcripts of a
handful of fifteen-year-old-and-older lectures he had given at Cambridge
University. New Millennium was mostly an audio publisher and had
apparently acquired the rights to publish the tapes of the lectures.
Unfortunately, the original contract—signed in 1988, at the height of A Brief
History of Time’s success—contained a clause allowing the publisher to
produce “in written form the text of said recording.” Al Zuckerman told a
reporter at the time that “maybe I was a little negligent in letting [the audio
publisher] put that language in there,” but added that he thought he was just
yielding permission to package a transcript of the audio along with the
recording—not to publish a separate book with Hawking as an unwilling
author.33
But that’s precisely what happened, and Hawking couldn’t stop it. He
tried complaining to the US Federal Trade Commission (FTC). “It was
professor Hawking’s concern that, if the Cambridge lectures were published
in book form, his large fan base would purchase the new book based solely
upon his name,” his lawyer wrote, “only to find that the work was not
authorized by Professor Hawking, that the quality of writing was not what
they would expect of him, and that its content was not new, but instead
merely a repackaged presentation of the material contained in ‘A Brief
History of Time.’”34
The FTC didn’t pursue the matter, and Hawking didn’t file a lawsuit. “I
just don’t have the financial resources and energy to take on this guy in
court,” Hawking’s lawyer recalled the physicist telling him. New Millennium
went ahead and published its book in Hawking’s name and made its money.35
The physicist, in the meantime, could only publish an outraged note on his
website:
Hawking’s name, like any other commodity, could be traded and used in
ways that the original owner never intended. Most of the time, though, the
physicist was cagey enough to get some benefit out of its use.
Wanting to live a purer, simpler life, they washed all their clothes by
hand and didn’t own a car and lit the house with candles in order to
avoid using any electricity.
It was all designed to give George a natural and improving
upbringing, free from toxins, additives, radiation, and other evil
phenomena. The only problem was that in getting rid of everything that
could possibly harm George, his parents had managed to do away with
lots of things that would also be fun for him. George’s parents might
enjoy going on environmental protest marches or grinding flour to
make their own bread, but George didn’t.42
George escapes the clutches of his Luddite parents with the help of a
(surprisingly short-tempered) scientist neighbor named Eric, his daughter
Annie, and their miracle-working computer.
The same year, Jane Hawking, Stephen’s ex-wife, published her memoir
of her time with Stephen, titled Travelling to Infinity. It was essentially a
lightly edited version of her earlier memoir, Music to Move the Stars. Some
of the harsher (and franker) passages had been softened or eliminated, but it
still painted a fairly dismal picture of life with the physicist. This was the
book that, seven years later, would become the Oscar-winning Theory of
Everything.
By this time in his life, Hawking had been the living embodiment of
scientific intellect for almost two decades. When he spoke on matters
touching upon science, his words were uniquely persuasive. Nobody else
could match the implicit authority about space, time, physics, or science in
general that Stephen Hawking carried. Such a strong voice attracts power.
And money.
It doesn’t take a lot of money to make a theoretical physicist very happy;
unlike experimentalists who need laboratories with large staffs and
expensive equipment, theorists can make do with a blackboard and a
graduate student or two. This makes them a cheap date for multimillionaires
looking to draw scientific superstars into their orbit. And over the years, a
number of wealthy people—mostly men—did just that, cultivating
relationships with leading theorists for reasons of their own.
Some were smart but lacked formal education, and relished in
surrounding themselves with the smartest people on the planet. In the 1970s
and 1980s, Werner Erhard, who made his money with a controversial self-
help program called est, not only forged ties with leading scientists around
the country but also hosted intimate conclaves where they gathered to discuss
topics of interest. (Hawking attended one of them; more on this in Chapter
12.) But at this point in Hawking’s life—the early 2000s—the new
undereducated billionaire on the scene was Jeffrey Epstein.
Epstein would later become infamous for his pedophilia and sex
trafficking. By then, a number of celebrities, including Prince Andrew of
England, lawyer Alan Dershowitz, and computer theorist Marvin Minsky,
stood accused of improprieties with Epstein-procured underage girls. (Both
Prince Andrew and Dershowitz have vigorously denied the accusations;
Minsky died in 2016.) In March 2006, a few months before Epstein’s first
indictment hit the headlines, he hosted a small conference on gravity in the
Virgin Islands. Hawking was there—and apparently on Epstein’s nearby
private island along with some of the other attendees of the conference: Kip
Thorne, Nobel laureate David Gross, and Harvard theorist Lisa Randall.
There’s also a picture of Hawking peering out the window of a submarine,
which Epstein reportedly had modified to allow Hawking to enter. Given the
date, Hawking was likely unaware of the allegations hanging over Epstein.
There is no evidence that Hawking did anything improper or, unlike a number
of other scientists, had any sort of contact with Epstein after the accusations
became public. For Epstein didn’t adopt Hawking in the way some other
billionaires had.43
In 2002, Hawking met oil baron George Mitchell and began a decade-
long relationship that ended only when the magnate died in 2013. “He’s
basically the guy who invented fracking,” says Hawking collaborator Andy
Strominger. “So he set up these two, three week retreats—he was a huge fan
of Stephen—which were basically designed to create an environment where
Stephen can work.” (Mitchell would go on to name an auditorium at Texas
A&M—his alma mater—after Hawking.) “George P. Mitchell was a
remarkable individual who combined vision with wisdom and persistence,”
Hawking said after the billionaire died. “Through sheer hard work and
dedication, he leaves behind an extraordinary legacy. It can be said of very
few people that they changed the world—but George Mitchell is among those
few.”44
Another billionaire that Hawking singled out for praise was Yuri Milner,
a Russian oligarch and Internet tycoon who helped invest hundreds of
millions of the Kremlin’s dollars into US corporations. “Yuri Milner is
something of a visionary,” Hawking enthused in 2016. “He sees that while
there are many good causes and pressing problems, ultimately our chances of
thriving as a species depend on tending and feeding the precious flame of
knowledge.”45
Milner met Hawking at a conference in Moscow in 1987—Milner has a
background in theoretical physics—and helped to tend and feed Hawking a
quarter century later. In 2013, Milner awarded Hawking a “special” version
of the billionaire’s new “Breakthrough Prize,” which came with a hefty $3
million purse. And Milner was, in fact, the “rich friend” who funded
Hawking’s 2016 trip to Harvard to work with Andy Strominger on soft hair.
“You know, it cost half a million to fly him there,” recalls Marika Taylor, one
of Hawking’s former students. “Yuri Milner gave him the half a million, so
here he had to believe that this was going to be a scientific breakthrough…
but in reality, it wasn’t a scientific breakthrough, right?”46
The relationship between the billionaire and the physicist got closer over
time. Hawking put his name to two of Milner’s hugely expensive and
somewhat daft scientific projects—a search for alien radio transmissions,
and an attempt to send ultra-tiny spacecraft to the nearest star—either of
which would likely have gotten little attention without a big name like
Hawking giving them his imprimatur.
“[Hawking] was very much conscious that if he wanted to keep in the
public eye, he needed sponsorship; you know, he needed support,” says
Taylor. And that support—and Hawking’s public profile—was to a large
degree dependent upon making headlines. “He was often asked to comment
about topics where he should have said, ‘You know, actually I don’t want to
comment about this; this is not something I have the expertise for,’” Taylor
says. “He would admit this privately, but he said it in public because… there
may have been sponsorship involved or he may have been doing it because
keeping him in the public eye helped the sponsorship.”
Though Milner’s projects were arguably a little kooky, they probably
appealed to Hawking on some level. It didn’t take much for Hawking to
swallow whatever discomfort he felt about straying from his expertise to put
his voice behind them. “Many of these things were not things he wanted to
talk about,” says Taylor. “But he felt he should.”47
For example, Hawking’s support of the Milner spacecraft project was
helped by the fact that Hawking was personally an advocate of spaceflight.
He had argued on numerous occasions that the long-term survival of the
species depends on our eventually being able to leave the planet. Of all of
Hawking’s out-of-left-field stances, it was this one that led to the most
benefit for Hawking. Including one of his most iconic moments, again,
sponsored by the largesse of a titan of industry.
In October 2006, Hawking met multimillionaire Peter Diamandis, a
California businessman who himself had a great interest in space and space
technologies. Six months earlier, Diamandis had signed a contract with
NASA to use the shuttle landing pad at Kennedy Space Center.48
Diamandis’ new company, the Zero Gravity Corporation, needed the
landing pad to perform zero-gravity flights with its Boeing 727. That is, the
jetliner would take off and climb to twenty-four thousand feet. And then, in a
set of stomach-churning maneuvers, the plane would suddenly lurch up and
zoom to thirty-four thousand feet, then tip over and plummet back down
again. The idea was to behave essentially like an object in freefall—if the
pilot did the maneuvers just right, the passengers would feel no gravity at all
for about thirty seconds at a time. For years, NASA used a similar plane to
train NASA astronauts; it was dubbed the “Vomit Comet” for all-too-obvious
reasons.
Diamandis offered Hawking a zero-gravity trip. “And he said, on the spot,
‘Absolutely, yes,’” Diamandis told an audience some time later. In April
2007, Hawking—with medical professionals on standby—took the plunge.
Flashbulbs popped as two handlers helped him float into the air. For many of
his fans, this was the only image that they had ever seen of the physicist out
of his wheelchair. Emaciated, his hands curled at unnatural angles, he was a
spectral figure—with a look of delight on his face.49
The headlines across the world about Hawking’s zero-gravity flight
certainly didn’t hurt the Zero Gravity Corporation’s prospects. And as it
happened, at the time of Hawking’s flight, the corporation had bid on a
contract to provide zero-gravity flight services for NASA—a contract that
the corporation won in January 2008—and that NASA later regretted.viii The
corporation’s performance was often terrible—at times, its pilots failed to
do the maneuvers properly more often than they did them right, meaning that
the flights were all but worthless—yet all the public ever knew of the
corporation, if they heard of it at all, was the Hawking flight. From the
physicist’s point of view, a free zero-gravity flight for a little PR was a no-
brainer. Upon landing, he gushed about the experience. “I could have gone on
and on,” he said. And then he exclaimed: “Space, here I come!”50
For Hawking had his eye on an even more ambitious target than a free
zero-gravity jaunt: a free trip to outer space. He had already acquired the
ticket.
In 2006, Richard Branson, billionaire owner of the Virgin corporation,
was famously trying to set up a service to launch tourists out of the
atmosphere and (hopefully) bring them back to Earth in one piece. At the
time, the service, Virgin Galactic, was scheduled to begin in 2009. When
Hawking appeared on a BBC radio show in November 2006, perhaps
emboldened by his encounter with Diamandis, he made a not-so-subtle
appeal. “My next goal is to go into space,” he told the host. “Maybe Richard
Branson will help me out.”
Branson, more than happy to oblige, promptly promised to give Hawking
a ride to outer space. “With delight I found myself with what I understand is
the only free ticket for a Virgin Galactic space flight that Richard has ever
handed out,” Hawking later said.51
But a promise is not a trip, and Hawking eagerly awaited the opportunity
to cash in that free ticket. Branson asked nothing in return. In the meantime,
the billionaire’s engineers spent years working out kinks and overcoming
delays while trying to get their spaceship into shape. Just when it looked like
Virgin Galactic was on a clear path to commercial flights, in October 2014, a
Virgin Galactic ship undergoing spaceflight testing broke up high over the
Mojave Desert. One of the two test pilots was killed. Branson’s endeavor,
already long delayed, was on the ropes.
After a year of investigations, redesigns, and retrenchments, Branson was
nearly ready to unveil the redesigned spaceship. So, in December 2015, the
Virgin team stopped by Cambridge University for an inspirational chat with
the professor. Hawking fondly recalled his appearance on the BBC show
when he expressed his desire to go into outer space. “[The host] asked me
whether I was worried by the prospect of death. I replied that as my death,
according to the medical profession, has been predicted many decades
earlier, it did not overly concern me, but that there were still a few things left
on my bucket list. Near the top of that list was the desire to experience space
for myself,” he said. Accidents will happen and people will die, but space
travel is not for the weak-hearted. Dulce et decorum est pro astra mori.52
When Branson unveiled the new prototype spaceship in February 2016,
Hawking was unable to travel to the event, but he regaled the crowd with a
recorded message: “We are entering a new space age, and I hope this will
help to create a new unity,” he declared, naming the new vessel. VSS Unity,
gleaming, sported a painted banner on its side bearing a gigantic image of
Hawking’s eye.
Though the physicist himself never got to ride, that rocket will always
bear Hawking’s name.
Hawking was known as a scientist of the first rank and one of the most
celebrated popularizers of science of our age. Yet in the last decade of his
life he did little science of note—certainly none considered to be of high
import—and very little science communication. He was neither scientist nor
communicator as much as he was a brand.
The authentic Hawking, the man who had devoted his life to physics, and
who had a passion to be understood not just by his peers but also by the
public, is barely visible behind the image, the commercial product that he
had become. It’s a vexing, almost paradoxical situation: Hawking’s celebrity
had almost completely obscured the very elements of Hawking that had made
him a celebrity in the first place. To find them, one must travel further
backward in time.
Footnotes
i Einstein never won for his relativity theory, but for his quantum work
regarding what happens when light hits metal objects.
ii Galfard, in fact, thinks that Hawking didn’t actually reverse himself: “My
true belief is that Stephen never believed in the information paradigm that
information was lost. He just wanted to find where it had gone.”
iii Kip Thorne put the blame on Hawking’s “not… terribly strong student,”
presumably Galfard. Leonard Mlodinow, Stephen Hawking: A Memoir of
Friendship and Physics (New York: Pantheon), 193.
iv King later won a Pulitzer Prize for his work on Thurgood Marshall. I have
not detected any patchwriting in the sections I know King is responsible for.
v Among the people he thanks is Leonard Mlodinow, who assisted Hawking
not with the little biographies that had the problems with patchwriting, but
with the overarching introduction to the book, which didn’t seem to have the
same issue. (Leonard Mlodinow, personal communication with author.) In
our discussion, I did not indicate to Mlodinow that there were patchwriting
issues, much less where the problems were, so his unprompted description of
the sections he contributed would seem to rule him out as a culprit.
vi In the interest of full transparency, I should reveal that I, too, had an
unpleasant fight over intellectual property that involved New Millennium.
vii Zuckerman was unable to recall precisely which work it was, but the
timing and the size of the advance imply that it was The Grand Design.
viii Even without the Hawking stunt, the Zero Gravity Corporation would
almost certainly have won the bid; of the four companies in the running, it
was the only one that owned its own plane.
PART II
IMPACT
“You’ll have in your possession absolute perfection upon this earth. No-
one is so rich he may vie with your wealth, if you have given the question its
due.”
He said: “I did not ask the question.”
“Alas that my eyes see you,” said the grief-laden maiden, “since you were
too daunted to ask the question! But you saw such great marvels there—to
think that you should have refrained from asking then! There you were in the
presence of the Grail.”
—WOLFRAM VON ESCHENBACH,
Parzival
CHAPTER 6
BOUNDARIES (1998–2003)
It was a time of revolution. In the late 1990s and early 2000s, a series of
astronomical observations—particularly of distant supernovae and of the
cosmic microwave background—were forcing cosmologists and gravitation
experts to revisit one of the basic assumptions about how the universe
behaved. For those observations were showing that spacetime was acting in
an unexpected way: it was as if there were some mysterious substance, “dark
energy,” that was slowly, inexorably stretching the fabric of the universe.
Once cosmologists understood what these observations meant, they would
not just be able to pinpoint when the universe began; they would also divine
how it will end.
Stephen Hawking was not a part of this revolution. Back in the 1970s and
1980s, his work on black holes and singularities and the very early universe
put him at the very center of the swirl of scientific activity in cosmology and
general relativity. By the late 1990s, he had found himself on the sidelines,
trying to stay relevant. Not just scientifically. A Brief History of Time was
more than a decade old, and he hadn’t published any major popular work
since. Hawking’s story of the universe was at risk of getting stale—or,
worse, he might lose his audience entirely.
As he turned fifty years old, Hawking’s struggle had taken on a new
dimension: he was no longer at the very center of physics or even of the
popularization of science. He would have to fight to avoid drifting to the
periphery.
The science of cosmology is largely about how the universe began and how
(whether!) it will end. Both of these questions, fundamentally, are about
geometry: about the shape of the cosmos. For, as Einstein’s theory shows, the
shape of the universe describes how it behaves not just through space, but
also through time. The fabric of spacetime encompasses everything in our
universe, no matter how far distant in space or in time it might be from you—
including the very start of the universe and the finish, if it ever happens. Just
as surely as the Fates of Greek mythology could predict any event by
examining the warp and weft of the fabric they wove, cosmologists can figure
out the origin and ultimate fate of the universe by understanding its shape on
the largest scale.
Just like any other geometric shape, the curve of spacetime looks rather
different on different scales, even though they’re all part of the same object.
The surface of the world that we live in is a good analogy. How should we
describe its shape? On the smallest scale—looking at, say, the surface right
beneath our feet—it’s generally pretty much flat. Zoom in far enough, and the
surface looks like a smooth, flat plane, even if you’re standing on a little
hillock or in a shallow pit. You can describe a location on such a plane with
two coordinates: call them x and y or N and E or something else, but two
numbers are generally sufficient to describe the location of any terrestrial
object. Also, with very good accuracy, on this small scale, you can figure out
the distance between any two points on the plane with the Euclidean metric
—the Pythagorean formula. In some sense, the world is flat; it’s a Euclidean
plane.
Things get a little more complex as we zoom out. The world still looks
more or less like a flat plane—a two-dimensional paper map does a pretty
good job at describing how far apart two buildings are. But the countryside
isn’t perfectly flat. Subterranean forces have warped the surface, creating
hills and valleys. A paper map of hilly San Francisco might not give you a
very good sense of how long it will take to walk from one building to another
across town; unlike a hike in nearly flat Kansas City, you have to add time to
the journey to account for all the scrambling up and down hills that aren’t
typically pictured in a flat map. On the middle scale, the flatness assumption
begins to break down; local features like hills and valleys can distort the
distance you would expect to travel based upon a typical two-dimensional
map.
Zoom out farther still, and you realize that the world isn’t really flat at all:
it’s round. You don’t notice this in everyday life; you typically feel that
you’re living in a flat(ish) world, interrupted only by hills and valleys. You
could spend your life blissfully unaware that you live on a ball rather than on
a flat pizza pie—if you’re stuck on the planet, it takes some relatively
sophisticated measurements, such as observations of visual distortions of tall
objects at a great distance, or differences in solar angles at widely separated
cities—to prove that we live on a sphere rather than a plane (or a donut or a
saddle or a bowl, for that matter). Even though on the smallest scale our
geometry is a flat Euclidean sheet, the topology of the world as a whole is a
sphere.i
The same sort of analysis holds for our universe, according to the theory
of grelativity; it’s just that we’re dealing with four-dimensional spacetime
instead of two- or three-dimensional space. On the smallest scale, our
universe is described not by the Euclidean metric of flat space, but by the
more complicated Lorentzian spacetime metric. That’s the fundamental shape
of the fabric of spacetime, the spacetime we move through on a daily basis.
Just as our little corner of the world appears fundamentally like a copy of
simple, smooth, flat Euclidean space, our little corner of spacetime appears
like a copy of Lorentzian spacetime. This space is “flat” in some sense: even
though it’s four dimensional and includes time as well as space, it isn’t
distorted by creases or pits (at least on the scales we encounter locally). This
is spacetime undisturbed by gravity or other distortions: the spacetime of
special relativity.
Zoom out a bit, and on the middle-level scale—say, looking at a solar
system or a galaxy as a whole—we can see that spacetime isn’t totally “flat.”
As the equations of general relativity predict, bundles of mass and energy,
like planets or stars or solar systems or galaxies, cause spacetime to curve.
For about a century, scientists have been able to spot that curvature by
watching a star’s light get pulled as it passes near the sun during a solar
eclipse, or by spotting the shapes of distant galaxies being distorted as their
light passes near huge lumps of mass. And with LIGO, scientists have just
begun to detect the ripples of gravitational waves disturbing the flatness of
our little neighborhood of spacetime. This is the spacetime of general
relativity: a dynamic fabric that’s still smooth but not quite flat, as it is
distorted by mass and energy and gravitational waves.
Then there’s the third level. Zoom out to consider the cosmos as a whole.
What is its topology? Is it flat like a (higher-dimensional) pizza? Round like
a ball? Or a donut or a saddle or a bowl? These are questions of cosmology:
among the biggest questions that cosmologists can possibly answer.
Because the shape of the universe encodes time as well as space,
knowing the topology of our manifold tells us about the universe of the
future.ii A sphere is a compact object—it doesn’t stretch in all directions for
infinity—and must be bounded. This means that if our universe is “spherical”
our cosmos will have a well-defined end. The universe will expand for a
while, all the galaxies and mass and energy flying apart as it does. The
expansion will slow down until the collective gravity forces everything to
start contracting again, stars and galaxies get pulled together again, and
everything finally ends in a “Big Crunch.” If, instead, our universe is flat like
a pizza, or warped like a saddle, it expands and expands without limits either
in space or in time. The universe keeps flying apart ad infinitum.iii
The equations of general relativity don’t tell us about the topology of the
universe; we have to gather data to figure it out. Just as it took smart
measurements and sophisticated reasoning for the ancients to understand that
the world was round, it took astronomers a tremendous amount of
observation and calculation to get a handle on what the overall shape of the
universe might be. And in the late 1990s, after many years of effort, they
finally began to succeed.
Two groups of astronomers were studying distant supernovae—stellar
explosions so bright they are visible half a universe away—when they each
noticed something very bizarre. The more distant supernovae were fainter
than expected, which seemed to imply that the expansion of the universe was
not slowing down fast enough to lead to a collapse; there was no way that the
universe was going to stop expanding and start contracting again into a Big
Crunch. It was an amazing discovery: the universe was expanding so fast that
it would have to expand ad infinitum. For the first time, cosmologists could
say something about the topology of the cosmos: it wouldn’t collapse like a
spherical cosmos would. This alone would have been worthy of a Nobel
Prize (which Saul Perlmutter, Brian Schmidt, and Adam Riess won in 2011).
But there was more to their discovery.
Not only was the expansion of the universe not slowing down very fast,
but it seemed to be speeding up. This was forbidden by the field equations of
general relativity—unless you added a little mathematical kludge known as a
cosmological constant. The cosmological constant, in some sense, is an
energy that’s a part of the fabric of spacetime itself, an energy that, like other
matter and energy, makes spacetime curve.iv But it has the opposite effect as
gravity; it pushes things apart rather than drawing them together. This
mysterious, antigravity-like dark energy contributes to the shape of the
cosmos—a shape that scientists were increasingly certain was infinite in
extent. The universe was not curved like a ball and primed for recollapse; it
had to be flat or saddle-shaped, and would expand forever and ever as
galaxies fly ever farther away from one another.
The two supernova-hunting teams published their results in 1998,
sparking a storm of excitement in the scientific community. The idea of an
accelerating universe, of a cosmological constant, of dark energy, was so
unexpected that it totally upset the pet theories of most of the cosmologists in
the world, Hawking included. For Hawking had spent many years pushing a
beloved theory, the no-boundary proposal, that featured prominently in A
Brief History of Time. Unfortunately for Hawking, the no-boundary proposal
seemed to imply that the universe would end in a Big Crunch—precisely the
result ruled out by the new supernova observations. The observations were
so compelling that Hawking had to admit that a Big Crunch was ruled out. So
in the 1999 Atlanta press conference, when Hawking said that it was “very
reasonable” that there should be a cosmological constant that would make the
cosmos “keep flying apart forever,” it seemed like he was finally giving up
on a theory that was one of the cornerstones of his career.
As with many things Hawking, the truth was quite a bit more complex.
One year earlier, in March 1998, Hawking gave a grand address at the
White House. President Bill Clinton, First Lady Hillary Clinton, and assorted
political and scientific luminaries were in attendance. After the lecture, the
audience peppered Hawking with questions. “Within the past month, we have
seen evidence suggesting a strong, repulsive force in the universe—an anti-
gravitational force causing the universe to expand, surprisingly, at an
accelerating rate,” Vice President Al Gore said. “How surprised were you
by this finding? What are its most important implications?”3
Hawking made his skepticism clear. “What the Vice President is referring
to is some observational evidence that suggests that there may be an anti-
gravitational force that would cause the universe to expand at an increasing
rate,” Hawking said. “The existence of such an anti-gravitational force is
very controversial. Einstein first suggested it might exist, but later regretted it
and said it was his greatest mistake. If it is there at all, it must be very
small.” He was still convinced that a Big Crunch was coming. “But don’t
worry, the Big Crunch won’t come for at least 20 billion years,” he said.
Even though it was nominally Hawking the scientist who addressed the
audience, it was Hawking the celebrity who got most of the attention. Hillary
Clinton read out a question submitted over the Internet. “How does it feel to
be compared to Einstein and Newton?”
“I think to compare me to Newton and Einstein is media hype,” Hawking
declared.
“I must say, you did look good at the card table,” Clinton responded.
The card table in question featured prominently in Hawking’s lecture
earlier in the evening. At the very start of the talk, the physicist played a short
clip from Star Trek: The Next Generation in which the physicist had been
conjured in the twenty-fourth century to engage in a battle of wits—poker—
with none other than Newton and Einstein. “All the quantum fluctuations in
the world will not change the cards in your hand,” Einstein declares. “I call.
You are bluffing, and you will lose.”4
“Wrong again, Albert,” Hawking gloats, a huge grin on his face, as a
motorized arm reveals his hand. Four of a kind.
Hawking may have considered it “media hype” to treat him as the
intellectual equal of Einstein and Newton, but it was an image that was
central to his public identity. In May 1999, viewers of the prime-time cartoon
The Simpsons were visited by “Stephen Hawking, the world’s smartest
man.” Armed with a gadget-filled flying wheelchair and an IQ of 280, the
physicist makes fools of the Springfield locals. But he does bond with one
character: the buffoonish, pastry-gobbling Homer Simpson. (“Your idea of a
donut-shaped universe is intriguing, Homer,” the animated Hawking says
over a round of beers at the local pub. “I may have to steal it.”v) Even in
Hawking’s appearances on TV commercials—such as his 1999 ad for
Specsavers opticians, which reportedly earned him £100,000—his words
were a conscious echo of Newton’s. “For me, physics is about seeing further,
better, deeper,” his voice declares. Onscreen, a CGI spaceship and a new
pair of glasses make the shoulders of giants entirely unnecessary for his
transcendent vision.5
The jacket copy on Hawking’s books did nothing to dispel the media hype
either. “Stephen Hawking is the Lucasian Professor of Mathematics at the
University of Cambridge and is regarded as one of the most brilliant
theoretical physicists since Einstein,” read the author biography in Universe
in a Nutshell, published in 2001. The Lucasian Professorship at Cambridge
is arguably the most prestigious academic post in the world—made famous
by one of its previous occupants, Sir Isaac Newton. Hawking was the heir
apparent to the legacy of Newton in more ways than one.
It was sometimes a heavy mantle for the physicist to wear. Moments after
dismissing the comparison to Newton and Einstein as “media hype,”
Hawking suggested another reason that he might be considered the world’s
smartest man. “I fit the popular stereotype of a mad scientist or a disabled
genius or, should I say, a physically challenged genius, to be politically
correct,” he said. “I am clearly physically challenged, but I don’t feel I am a
genius like Newton and Einstein.”6
Nor did his colleagues put him in that category. In 1999, the magazine
Physics World asked some 250 physicists around the world to name the five
physicists, living or dead, who had made the most important contributions to
physics. Einstein topped the list with 119 votes, followed by Newton, and
James Clerk Maxwell with 67 votes. Galileo was sixth, Richard Feynman
seventh, Paul Dirac eighth. Gerard ’t Hooft and Stephen Weinberg, who each
helped understand the weak force, and Charlie Townes, inventor of the laser,
got two votes each. Hawking—along with a couple of dozen others, such as
Martin Rees and John Wheeler—wound up at the bottom of the list, each
with a single vote. It was an honor merely to be included, but the poll left
little doubt that, generally, physicists didn’t think his work to be in the same
league as a Newton or an Einstein or even a Dirac, no matter what
impression the mass media gave.7
Hawking knew that his disability deeply affected the way the public
perceived him, transforming him from a mere human being into a living
metaphor. The image of a man trapped in a quadriplegic body soaring,
intellectually, to the ends of space and time; the idea of a person who is
unable to speak, who can only, with heartbreaking amounts of labor, struggle
to utter a few words a minute, becoming one of the world’s best-selling
communicators of science—the physicist’s story was powerful and deeply
ironic. And Hawking was keenly aware of that irony.
“He loves that Hamlet line,” says director Errol Morris, “the ‘nutshell’
line. And why wouldn’t he?”8
In Act II of Shakespeare’s play, a courtier suggests to Hamlet that
Denmark is too small for him, and restricts his mind. The doomed prince
responds, “O God, I could be bounded in a nutshell and count myself a king
of infinite space, were it not that I have bad dreams.” Says Morris: “And in
the Shakespeare itself, there’s a very strong irony, which he is playing on,
clearly.” The transcendent idea, the beautiful image of destroying all
limitations with the power of the mind, is snatched away by attendant
nightmares. Hawking was stoic—and closed-lipped—about whatever bad
dreams pulled him back to uncomfortable reality.
The Universe in a Nutshell was Stephen Hawking’s first major book since
his groundbreaking best seller, A Brief History of Time, published more than
a decade prior. It was the first serious literary attempt to capitalize on his
stature as the reigning monarch of physics popularization. But in the
intervening time, there had emerged some new pretenders to the throne. Most
notable was Brian Greene, a theorist at Columbia who studied hypothetical
subatomic objects known as superstrings. Greene’s book, The Elegant
Universe, sold at a steady clip since it was published to rave reviews in
1999, and wound up on the paperback best-seller lists for several months the
following year. The market for science nonfiction books was still there; the
question was just whether Hawking could capture it once again.
Hawking lamented that many of his readers “got stuck” in the beginning
chapters of Brief History and never finished the book.9 Nutshell, unlike Brief
History, would be lavishly illustrated, all the better to make it easier to
understand and reach a broader audience. However, it would be a struggle
for Hawking to produce new material for the book. This time, he would have
help from the get-go. It would come not just from a graduate student (Thomas
Hertog) but also from science writer and erstwhile Hawking biographer
Kitty Ferguson who recalls:
Though the raw material was a total mess, it would have been “unthinkable”
for Harris to turn down a proposal from Hawking. Ferguson did her best to
even things out, but in the end, even the heart of the book—the illustrations—
wound up being somewhat haphazard. One of Hawking’s assistants, Neel
Shearer, was asked to generate an idea for an image in the book. Shearer
created a Monty-Pythonesque mockup of God shooting a lightning bolt and
creating an explosion in space, and turned it in. “I hadn’t appreciated that it
was an image that was going to be published in the book,” he says. “I
imagined that it was an image that was going to be passed to the publisher’s
illustrator to be made into something proper.” It was a great surprise to
Shearer when he saw the completed book and his original illustration was in
there.11
When Universe in a Nutshell came out in November 2001, it immediately
shot up the best-seller lists, rapidly reaching number 4 in the New York
Times. Reviews were mixed. Readers praised the beautiful illustrations, but,
as one reviewer complained, “your £20 buys you a lot of repetition of things
you first read in A Brief History.… [T]he images make this book a more
handsome object, but also cloak the brevity of the text, a scant 100 pages of
unadorned print. That means that quite a lot of things which should be
explained are skimped.”12
There was a good deal of overlap with Brief History, but there were
definitely new elements, such as an explanation of the holographic principle,
an increasingly important area of exploration in cosmology and black-hole
physics. And some reviewers praised his exposition—fellow physicist Joe
Silk called the book “a delight to read” because of “Hawking’s caustic
asides and his infallible optimism.” But many of the reviews were lukewarm
or harsh, such as one written by physicist and science popularizer John
Gribbin—who himself wrote a biography of Hawking.13
Whereas previously the party line was that A Brief History was all his
own work (which anyone who has seen the early draft knows cannot
be true),vi this time Hawking thanks Ann Harris and Kitty Ferguson,
“who edited the manuscript.” I don’t know the work of Ann Harris, but
some of the less technical parts of the present book certainly read like
the work of Kitty Ferguson.…
Somebody at Bantam should have had the guts to tell Hawking that
his jokes aren’t funny, and Hawking, assuming he hasn’t begun to
believe his own publicity, should have told them in no uncertain terms
to tone down the blurb. “Great” is an adjective that should be used
sparingly, and when used in science reserved for the likes of Albert
Einstein and Richard Feynman.14
Stephen was not the only one who had been busy writing. Shortly after her
marriage to Stephen had fallen apart in 1990, Jane had decided to embark
upon a “new project, a project of my own, through which I could prove to
myself, if to nobody else, that I had a brain which was both capable and
inventive,” she wrote. She wanted to write her memoirs, not about her life
with Stephen, but about her experiences as an Englishwoman setting up a
second home on the Continent. “It would consist of amusing anecdotes and
practical information, aimed at the considerable market of British buyers of
homes in France.”17
However, for some reason, her literary agent at the time was unable to
generate much interest. One publishing house made an offer to print the book
if Jane would commit to writing her memoirs as well. Insulted, Jane refused,
and after feuding with her literary agent, resolved “to publish my French
book myself, whatever the cost, and to deprive him in perpetuity of any
commission on any other book that I might write.” Once her contract with her
agent expired in 1994, Jane self-published At Home in France, and in 1995,
shortly after her divorce from Stephen was finalized, she accepted an
invitation by Macmillan to write her memoirs, her tale of a “quarter of a
century of living on the edge of a black hole,” married to Stephen.18
The result, Music to Move the Stars, hit the bookshelves in 1999 to a
flurry of press attention. One reviewer called it “a small book about a small
life, bounded by the all-consuming needs of another, occasionally tinted with
martyrdom and alarmingly empty of dreams.” But the physicist himself was
no longer the sole author of the narrative of his life. The woman he once
loved—and who had devoted herself to his care—was painting him as a
ruthless and selfish tyrant, “an unruly, demanding, and assertive child who
needed my protection both on account of his physical helplessness and that
peculiar naivety, born of his hyper-intelligence, which can blind the famous
to the Machiavellian subtleties of personalities and motivations.”19
For Jane, a central narrative of her marriage—and much of the reason for
its dissolution—rests upon a central irony: “Her Christianity gave her
strength to support her husband, the most profound atheist,” as one journalist
put it. Jane had inherited a deep faith in God from her mother “which had
sustained her through the war, through the terminal illness of her beloved
father and through my own father’s bouts of black depression.” Stephen’s
atheism offered only a “bleakly negative influence which could offer no
explanations, no consolation, no comfort and no hope for the human
condition,” and, despite her faith, pulled her away from God. It wasn’t until
after Stephen left her, Jane wrote, that she “became convinced that God does
exist as the ultimate power of goodness.” For Jane, the separation was
almost preordained by the Almighty, who had recognized that she was in the
wrong marriage and would not allow her to find true happiness until she
found the right helpmeet.20 vii
Jane’s version of events struck a nerve—it was a classic narrative trope
—and it soon provided fuel for a production at the Theatre Royal in Bath in
late 2000. The play, God and Stephen Hawking, was no Doctor Faustus,
and the dramatist, Robin Hawdon, was no Marlowe. He was, however, a
veteran playwright who had written a wildly successful comedy, The Mating
Game. And while Stephen had refused to respond to Hawdon’s attempt to
engage him in the project, Jane had made some “quite helpful suggestions.” It
was precisely the sort of thing to get the literati talking Hawking.21
God and Stephen Hawking pitted the two titular characters in a battle of
wits—all about whether the almighty truly exists or was a figment of human
imagination. Jane was a divine instrument, a “great weapon” that God used to
try to win the battle that Stephen wasn’t even aware he was engaged in:
JANE: Perilous?
Both Jane and Stephen are tempted to betray each other, but only Stephen
succumbs, running off with his nurse, Elaine. After pronouncing her closing
line, “All the same, whether you want him or not, God be with you,” Jane
leaves, never to be heard again, ceding the stage to Stephen and his creator to
battle it out amongst themselves.22
The repetitive squabbling over religion is both inadequately written and
ultimately fruitless,” complained one reviewer. Another argued that the
“extraordinarily misconceived” play “subjects Hawking to the humiliation of
taking part in a rigged, bogus, clumsily written debate.” Stephen himself
found the play “deeply offensive and an invasion of my privacy,” he told
Physics World. “I could probably have got a court order against it, but I have
never approved of public figures using [threats] of legal action and it might
have just attracted more attention to a stupid and worthless play.” Martin
Birkinshaw, a cosmologist who helped Hawdon with the science in the play,
didn’t see it that way. “I rather feel that it is up to the playwright to decide
who to write about. Stephen has been in the public eye for so long, and
through such a wide range of television programmes and books, that surely
this play can make little difference?”23
It was unusual for Hawking to complain about how he was portrayed. For
example, in 2000, a website with the stylings of gangsta science rapper “MC
Hawking”—complete with synthesized voice fronting hip-hop beats—went
viral. In response to lyrics such as:
Hawking’s assistant sent the author an email stating that the physicist was
“flattered.” But as crude as MC Hawking’s lyrics were, the tongue-in-cheek
portrayal of Hawking as a badass gangsta for science didn’t conflict with the
image that Hawking had been enjoying in popular culture. It was another
thing entirely to paint him as needy, arrogant, manipulative, or—worse—
vulnerable. That was not a narrative of his choosing. Nor was it one that he
could control.25
The counternarrative gained strength in 2000, when the first reports began
to emerge that Stephen had sustained mysterious injuries—and that the police
had begun an investigation into the circumstances. Stephen and Elaine
maintained their silence. But there were dark rumors that got even worse just
before his sixtieth birthday in 2002, with the incident in which he broke his
hip. Though Hawking insisted that he had run into a wall, and his assistant
affirmed that the accident was because of excessive haste (“He was late for a
meeting and running on Hawking time, as ever”), the questions about abuse
rumbled just underneath the surface. It accompanied him and Elaine to
scientific conferences and public lectures. And it brought not just Hawking’s
physical dependence on others into the public arena, but also his sexuality.26
The sexual elements that contributed to the end of his first marriage, and
made their appearance in Jane’s tell-all memoir, had brought Stephen’s
libido into the sphere of public discourse. There was no longer any way to
maintain the image of the physicist as the ultimate force of rationality, a being
who, as sociologist Hélène Mialet put it, “can do nothing but ‘sit there and
think about the mysteries of the Universe,’ this intellect liberated from his
body and seemingly emancipated from everything that clutters the mundane
mind.” Stephen Hawking was a flawed human being with an ego and a libido
and who could be snickered about behind his back. It was about this time that
the press started getting interested in Hawking’s habit of visiting nudie bars,
a habit he might have picked up in the States.27
“At Caltech, there was this tradition of going to strip clubs, I believe
quite a bit,” says physicist Neil Turok, who collaborated with Hawking a
number of times before going on to run the Perimeter Institute of Theoretical
Physics in Waterloo, Canada. “And Stephen, when he was here [in
Waterloo], his assistant says to me one day, ‘Neil, do you know if there are
any gentlemen’s clubs?’”28
“And I said, ‘What do you mean gentlemen’s clubs? I have no idea what
you mean.’ So she said, ‘Gentlemen’s clubs.’ And I said, ‘You mean like in
London, posh clubs, you know, for upper-class people like in Mayfair?’ I
literally had no idea what she was talking about. And she finally says,
‘Strippers!’” Turok’s tone lowers, and he speaks between clenched teeth.
“And I said, ‘If you dare take Stephen to a strip club in Waterloo, I will
explode! This, it will ruin our name as a scientific… we cannot… So don’t
take him there. Even if there is one, don’t take him.’”
Because Hawking was so dependent on others for his mobility, this yen
for strip clubs (and even more risqué entertainment) naturally had the
potential to put his hosts, his nurses, and his students in an awkward position.
The press first noticed Hawking’s predilection for nude bars in 2003,
when Hawking was spotted with actor Colin Farrell in Stringfellow’s, a
London strip club. According to Peter Stringfellow, the club’s owner, he
spent more than five hours with a stripper named Tiger—but he didn’t reveal
what the couple did or discussed. Ever after, Stephen’s visits to gentlemen’s
clubs and eventually sex clubs—along with the nitty-gritty of his spousal
relations—were grist for the tabloids, which seemed to delight merely in the
image of a nearly immobile physicist in a wheelchair becoming sexually
aroused.29
Despite the best efforts of the tabloids, the prurient stories never reached
the vast majority of Hawking fans. He stayed almost entirely aloof; his image
was still of the same Star Trek–worthy intellect, the same symbol of
rationality, the same stoic hero who, uncomplaining, overcame the most
profound adversity to understand the workings of the cosmos in a way no
other mortal could.
It was a time when the workings of the cosmos were becoming more visible
by the day. In 2000, the supernova observations were still new, and
cosmologists were just beginning to grapple with the possibility of dark
energy, a totally unexpected substance that stretches and curves the fabric of
spacetime. And there was another set of observations just around the corner
—observations that had the potential to be just as revolutionary for
cosmologists’ understanding of the shape, origin, and fate of the universe. In
retrospect, Hawking would later say that the ongoing discoveries were the
most exciting developments in physics during his career.30
Flash back thirty-five years, to 1965, when two engineers at Bell Labs in
New Jersey—Arno Penzias and Robert Wilson—were having trouble with
an antenna that was designed to pick up microwaves designed for satellite
communications. Wherever they pointed it in the sky, the antenna hissed with
noise. It seemed like the equipment was malfunctioning; no matter how they
tinkered with it, no matter how carefully they controlled where they pointed
the antenna, they couldn’t seem to eliminate the noisy static.
At the same time, a few miles away, at the Princeton campus, physicists
Bob Dicke and Jim Peebles were doing some calculations having to do with
the origin of the universe. If the universe was really born in a Big Bang—a
sudden explosion that created the manifold of spacetime all at once—the tiny,
early universe must have been very, very hot and glowed brightly with highly
energetic light. But as the cosmos expanded, it cooled down, and the light—
which was everywhere in the universe—stretched out. As light stretches, it
becomes less and less energetic, changing from gamma rays to X-rays to
ultraviolet to visible light to infrared to microwaves. So, the Princeton
scientists reasoned, billions of years after the Big Bang, that remnant light
should be visible as microwaves coming from all directions in the sky.viii A
microwave antenna pointed upward should be able to detect that light—
which would be a background hiss that wouldn’t go away no matter where
the antenna looked. They were planning to build a microwave antenna to see
if they could spot that hiss when they got a call from the engineers at Bell
Labs asking for help with an odd little static problem they were having.
This was the discovery of the cosmic microwave background (CMB)
radiation. It’s hard to explain this sort of omnidirectional, almost uniform
hiss of microwaves unless you assume that the manifold of spacetime was
once very small and very hot, and expanded to give us our present-day
universe. Penzias and Wilson received the Nobel Prize for their discovery;
Dicke and Peebles, the theorists, got a nice thank-you.
But there was another layer of discovery just out of reach of 1970s
technology. Even though the microwave static looked the same in every
direction, the theory says that it can’t be quite uniform; as scientists like to
say, it was anisotropic. That is, if you look hard enough, you should be able
to see differences in the microwaves at different scales, hot patches that had
more energetic light than average and cold regions that had less energetic
light. The fabric of spacetime in the early universe was rippling with energy,
squashing and stretching out clumps of matter at various times. This, in turn,
made the spacetime of the early universe curve on various different scales. If
the Big Bang model were correct, those ripples and curves in the early
universe should have caused fluctuations in the energy of the CMB, hot spots
and cold regions of various sizes. Penzias and Wilson’s antenna wasn’t good
enough to see those variations. But a better instrument should be.
That instrument was the Cosmic Background Explorer (COBE), a satellite
launched in 1989, a quarter century after the discovery of the CMB. The
COBE team quickly proved that the CMB did, indeed, have hot and cold
regions, just as the theory had predicted. This discovery, too, won a Nobel
Prize. However, COBE’s instruments weren’t good enough to resolve those
spots with very much detail. It was like a nearsighted person looking at an
eye chart without glasses; it’s not hard to recognize that there’s something
written on the chart, and perhaps even to read the largest few letters at the
top. But the chart is a blur, and it’s impossible to gain enough information to
figure out what the smaller letters are saying. COBE was looking at the CMB
with blurry vision, unable to resolve, or extract information from, smaller
spots in the radiation. And encoded in the size of those smaller spots was
nothing less than the age of our universe—and its shape.
The theory that describes the hot and cold spots caused by the rumbles of
energy in the early universe also implies that there should be a characteristic
spot size—a maximum breadth for the hot or cold spots caused by the
rumbles in the early universe.ix These maximal spots were predicted to be
about a degree across in the sky. And this gave cosmologists a way to
measure the curvature of the universe.
If we lived in a spherical universe, distant objects would seem large in
the sky, magnified just as if we were seeing them reflected in a bowl-shaped
shaving mirror. If we lived in a saddle-shaped universe, the opposite would
be true. Only if the universe were flat would those distant objects be
undistorted, their apparent sizes in the sky neither magnified nor reduced by
the curvature of spacetime.
All scientists needed to do to pin down the curvature of the universe was
to measure the size of those maximal hot and cold spots in the CMB. If they
were larger than expected, it was evidence of a spherical, closed universe. If
they were smaller than expected, it meant that we were in a saddle-shaped
one. And if they were precisely the size theorists said they would be, well,
that would be a very strong indicator that our universe was flat. It was a
measurement cosmologists had been hoping to perform for decades—but
microwave antennas weren’t sophisticated enough to get the smaller-than-
one-degree resolution needed for the requisite comparison.
Flash-forward again to 2000. After years of effort, scientists had finally
managed to improve their microwave detectors enough to start resolving hot
and cold spots about a quarter of a degree across. Data began to trickle out,
and as best as cosmologists could tell, those hot and cold spots were
precisely the size that theory predicted.
The universe seemed to be flat after all.
It would expand forever. There would be no Big Crunch.
If you find this hard to understand, you’re not alone. The Hawking-Hartle
no-boundary proposal was controversial from the start; the shift from regular
time to imaginary time made it hard to interpret, and it wasn’t at all clear that
the mathematical calculations Hawking and Hartle were doing to describe
their picture were valid. As physicist and former Hawking student Don Page
wrote, “although at first sight the proposal looks conceptually clear (at least
to those who understand the concepts, so long as they do not worry about
details), when one looks at the details one finds that the proposal is not yet
mathematically precise.” This made it hard for many physicists to engage
seriously with the concept. “The idea has never been accepted,” says
physicist Neil Turok, who helped Hawking explore the proposal in the
1990s. “I would say 90% of cosmologists or theoretical physicists don’t
even form an opinion. Of those who do, 90% of them would say they
probably don’t agree with it, or they thought there was some problem with
it.” Nearly four decades after Hawking first lofted the no-boundary proposal,
physicists are still fighting over its validity.32
While this idea might seem extremely complex to some, to others, there’s
a certain elegance, a certain beauty to the way Hawking dodged the problems
inherent in asking about the origin of the universe. And Hawking, like many
other theoretical physicists, believed that the true laws of the universe
possess a certain spare beauty. As a consequence, he became very attached
to the no-boundary proposal. “That just rang true to him. Whether that was
for cultural or religious or whatever reason, I couldn’t say, but I think he
thought it was true just because it’s simple,” Turok says.
The no-boundary proposal featured prominently in A Brief History of
Time, but when Hawking wrote the book in the late 1980s, the proposal
required that the imaginary-time-universe be rather symmetric—the Big Bang
at the beginning of the universe had to be matched by an equal and opposite
Big Crunch. And because our universe would end in a Big Crunch, it couldn’t
be flat. Moreover, for technical reasons, the no-boundary proposal seemed to
imply that there was no cosmological constant—that it is “not necessarily
zero, but zero is by far the most probable value.”33
And so things remained until the supernova data in 1998, the high-
precision CMB data in 2000, and, finally, the results from the successor to
COBE, the WMAP satellite, which were released in 2003. We finally had a
definitive answer; the universe was flat or very nearly so, and it would keep
on expanding forever. A Big Crunch was no longer a possibility. The cosmos
was open, not closed. And there was a significant cosmological constant.
As theories are wont to do when subjected to experimental pressure, the
no-boundary proposal began to wiggle and transform. In 1998, Neil Turok,
who was then a professor at Cambridge, suggested to Hawking that, hidden
inside his sphere-like, finite, imaginary-time universe, there might reside an
open, infinite, real-time universe rather than only a closed one that would
collapse in a Big Crunch. “It’s such a surprising thing. If you start with a
finite universe, inside it, you can create an infinite universe, so that’s very,
very paradoxical,” Turok explains. “Inside a small piece, you can form this
infinite spatial universe which lasts forever.” No longer was a Big Crunch an
absolute prediction of the no-boundary proposal. But there was another
problem.34
“The truth is, his proposal wasn’t really working. It predicted an empty
universe; you can have this picture of the universe coming from nothing, but
the prediction is that it’s empty,” says Turok. Even after tweaking the model
to force it to produce at least one galaxy, the model produced exactly that—a
universe with just one galaxy. Nothing more. And nothing that in any way
resembled our own universe. “We predicted an empty universe or one
galaxy; it’s not very good. Stephen said, ‘You know, don’t worry about that
too much. It’s still progress….’ So he was pretty happy-go-lucky about it. I
was actually amazed.”
Hawking was able to keep holding onto his proposal even as some of its
predictions failed and the mathematics frayed around the edges. The details
simply weren’t working out. But the details weren’t really the point in the
first place. In the 1980s, Hawking had presented a beautiful idea—a seeming
way to avoid many of the messy questions about how the universe came to
be. To Hawking, the no-boundary proposal had the ring of Truth with a
capital T, and that was enough, even if he and his collaborators couldn’t
work the kinks out of the mathematics. It just felt correct to him, and to
Hawking, instinct was a powerful driver. The big beautiful idea was the
important thing, not the details.
And Hawking had become less and less able to work out the details
himself. As he himself admitted, since he’d lost the use of his hands in the
early 1970s, he couldn’t do mathematics in the same way everyone else did.
He couldn’t write complicated formulae, doodle diagrams, or even store a
transient thought in an efficient way. That made it hard for him to manipulate
a lot of the necessary mathematical formalisms or to build the equations that
breathe concrete details into a beautiful and innovative, but skeletal, idea. “It
is difficult to handle complicated equations in my head,” Hawking told the
audience at the White House Millennium Lecture in 1999. “I therefore avoid
problems with a lot of equations or translate them into problems in geometry.
I can then picture them in my mind.”
“He has gradually trained his mind to think in a manner different from the
minds of other physicists: He thinks in new types of intuitive mental pictures
and mental equations, that, for him, have replaced paper-and-pen drawings
and written equations,” Kip Thorne wrote of his longtime friend. “Hawking’s
mental pictures and mental equations have turned out to be more powerful,
for some kinds of problems, than the old paper-and-pen ones, and less
powerful for others.” But, according to Andy Strominger, Hawking’s
disability definitely limited his reach as a physicist. “He did, of course,
develop amazing abilities to do things in his head that anybody else would
have needed a paper and pencil for. But even so, he couldn’t do as much in
his head as other people could do with a paper and pencil,” Strominger says.
“And, you know, and it got worse as the years went by and whenever it
would get to a detailed calculation, you know, he would go as far as he could
in his head, and then, you know, the rest would be up to his collaborators.”35
“Working with Stephen doesn’t mean working with equations. It means
working with words and with concepts. That’s how he leads my research,”
Hawking’s graduate student Christophe Galfard told a BBC film crew
making a documentary about Hawking in the early 2000s. Galfard was
helping Hawking not with the no-boundary proposal, but with the black-hole
information paradox—the question that led to Hawking conceding his bet
with John Preskill in 2004.36
Quantum theorists and string theorists, unlike general relativity specialists
—and unlike Stephen Hawking—had always been fairly confident that
information wasn’t permanently lost to a black hole, and by the mid-1990s,
Hawking was losing the argument. And then, in 1996 and 1997, a rash of
papers by physicists Juan Maldacena, Ed Witten, Andy Strominger, Cumrun
Vafa, and others seemed to settle the question fairly decisively against him—
or so the quantum theorists believed. “Beyond a shadow of a doubt…
information would never be lost behind a black hole horizon,” wrote
Stanford physicist Leonard Susskind in his book about his “war” with
Hawking over the information paradox. “The string theorists could
understand this immediately: the relativists would take longer.… Although
the Black Hole War should have come to an end in early 1998, Stephen
Hawking was like one of those unfortunate soldiers who wander in the jungle
for years, not knowing that the hostilities had ended.” Yet the Black Hole War
continued to claim casualties.37
“Andrew Farley turned out to be one of the two best graduate students that
I have ever had,” writes Peter D’Eath, one of Hawking’s early PhD students
who joined the physics faculty at Cambridge shortly thereafter. Under
D’Eath’s tutelage, Farley wanted to examine certain quantum-mechanical
effects after the collapse of a black hole: effects that strongly implied that
information would not be lost when falling past the event horizon.38
The department’s rules were that a student couldn’t formally enter the
PhD program until the fourth semester, when two assessors reviewed his or
her work and plans and gave the go-ahead for the student to register. Unless
the student were performing poorly, the assessment was mostly a rubber
stamp—with the added benefit of having experts give the student some
feedback on potential weaknesses and point out avenues for further
exploration. D’Eath, quite naturally, picked Hawking to be one of Farley’s
assessors. “Of course, in retrospect, this caused a serious amount of trouble,”
D’Eath notes. It was unexpected: D’Eath had a collegial relationship with his
former mentor, and frequently invited him and Elaine over for meals.
However, after Hawking began looking at Farley’s work, he no longer
seemed willing to visit the D’Eaths. “For the first time, Stephen had either
declined or put off any acceptance indefinitely.”
The assessment of Farley’s work was delayed for several months, and
there were some very odd changes of the rules—for example, the assessment
wound up using three rather than two assessors, and D’Eath himself was
“interrogated” about Farley’s work in May 1999. (“Indeed, I have never
heard of a supervisor being required to take part in such a process,” D’Eath
writes.) And when the assessment finally happened, it was clear that
Hawking was very displeased with Farley’s project. Hawking went on the
attack and eventually recommended to the university that Farley be kicked out
of the program.
Though the head of the department overruled Hawking, and Farley was
allowed to continue with his studies, that wasn’t the end of the matter. D’Eath
says that he later discovered that “Stephen had continued to try to have
Andrew Farley de-registered via the Degree Committee of Mathematics, then
the Board of Graduate Studies, then the Vice-Chancellor himself.… Each
time, Stephen was acting ultra vires, and each time his application was
denied.”
Farley himself didn’t elaborate on his adviser’s account, writing, “I have
nothing material to add other than Stephen’s unreasonable behaviour knocked
my confidence thereafter. Stephen was someone I had greatly admired as a
boy and up to the point of meeting him. They say never meet your heroes; this
is particularly apt.”
It is highly unusual in academia for such a powerful professor to marshal
all of the forces at his disposal to try to destroy a mere graduate student. It’s
even more unusual for the student to prevail. Though Farley went on to
complete his PhD, it was a Pyrrhic victory. He was never able to get a
postdoctoral position, and that meant that any hopes of an academic career
were dashed. Farley currently works as a compliance officer for a finance
firm. “However, I never gave up doing my own theoretical physics research
after Cambridge,” he writes. “I’m currently doing my own research on
gravitational lensing.… Ideally, I would now be doing theoretical physics
research full time as this is my passion.”39
Also dashed was Hawking’s quarter-century-old friendship with D’Eath.
The contentious “interrogation” about Farley’s work in May 1999 was the
last time that Hawking and D’Eath spoke to each other.
Ironically, Hawking was beginning to harbor doubts in his own mind
about information loss; the string-theoretic work coming out at that time was
powerful and convincing. Even though that sort of research wasn’t quite in
Hawking’s wheelhouse, it shook his confidence.
In 2002, Hawking assigned his student Christophe Galfard the task of
understanding one of the important Maldacena papers that had come out the
previous year. “Stephen asked me to have a look at that paper so I took a
little while to read it. A little while being about a year and a half,” Galfard
told a BBC film crew. The paper used some of the powerful new insights
coming out of the theoretical community to imply that Hawking was wrong
about the paradox. But the paper held firm despite their best efforts, with
Hawking providing the big ideas via oracular guidance and Galfard working
out the fine details. The documentary filmmaker told sociologist Hélène
Mialet that, during the course of filming, “it became quite obvious early on
that effectively, you know, his graduate students do all the actual work… in
terms of hashing the numbers and working through the equations and all of
that, and then they bring it to him, and he obviously assesses it and then
points them in different directions.” Working together in this way, Galfard
and Hawking dissected the Maldacena paper. “He thought that there was
something in there that was new and different,” Galfard says. “I do strongly
believe that it coincided with some deep ideas that he had for a long time,
and this gave him the mathematical framework to actually verify whether or
not the information paradox really was a paradox.” And as they ventured
deeper into the new territory, the two lay the groundwork for Hawking’s
Dublin announcement, his formal U-turn on the question of whether
information is lost in a black hole.40
But before the two could complete their work, Hawking was struck down
with a bad case of pneumonia; it was very grave, and Hawking was placed
on life support. “We didn’t know whether we would see him again. We were
very, very concerned,” Galfard said. It was three months before he was able
to leave the hospital.41
Hawking never expected to survive long enough to see any of his theories
falsified; he had already outlived his doctors’ dire predictions by an almost
unbelievable margin. But as his students and collaborators were keenly
aware, Hawking spent most of his years teetering on a stark boundary. His
emaciated frame seemed too fragile to contain a life within. Yet despite the
way his condition consumed his being, he was always stoic about it—he
even turned it into a source of fun. BBC reporter Pallab Ghosh wrote about
one encounter he had with the physicist in 2004:
The camera operator I was with wanted to make a last minute
adjustment to his lighting and so he asked Prof Hawking’s staff if he
could pull out one of the plugs in the office so that he could use the
socket for his equipment.
Without waiting for a response he pulled the plug and the room
was filled with a deafening siren.
Prof Hawking then slouched forward and I feared that my
colleague had inadvertently unplugged a vital piece of life-support
equipment.
Fortunately, it was the alarm to the uninterruptable power supply
to his office computer and he was slouched forward with mirth at our
incompetence.42
Footnotes
i This is really the essence of a “manifold”: it’s an object that might (or might
not) have a complicated topology, but if you look at it on the smallest scales,
it has a regular, smooth geometry.
ii And, of course, the universe of the past, though that’s a bit more
complicated. More on this shortly.
iii Technically, the relationship between topology, the fate of the universe,
and whether the cosmos is infinite or finite in extent in space and/or in time
is a bit more complicated than this, once you allow for a cosmological
constant, or if you don’t rule out some of the more exotic topologies. So this
is a bit of an oversimplification, even though the general principle holds.
iv Because this is a book primarily about Hawking and not cosmology, this
description merely skims the surface of just how weird—and how
revolutionary—this discovery was. For details about the supernova
observations, the cosmological constant, the cosmic microwave background,
and the implications for cosmology, see my Alpha and Omega, chaps. 4 and
5.
v Hawking would appear on The Simpsons several more times. He also
played a recurring role a number of years later in the geeky sitcom The Big
Bang Theory.
vi In fact, Gribbin had seen an earlier draft. More on this later.
vii Though it was clearly a source of tension, religion does not figure
prominently in what little Stephen wrote about his marriage to Jane or its
failure. He implied, laconically, that the breakup was due to another cause
altogether: “I became more and more unhappy about the increasingly close
relationship between Jane and Jonathan [Hellyer Jones],” a local choirmaster
who had taken up residence with the Hawkings. “In the end, I could stand the
situation no longer.” Stephen Hawking, My Brief History (London: Bantam
Books, 2018), 87.
viii This is a bit of an oversimplification. The remnant light is actually from
a time roughly 380,000 years after the Big Bang, when clouds of electrons
and protons—which were opaque to light—cooled down enough to form
hydrogen gas.
ix Because the CMB was released roughly 380,000 years after the Big Bang,
the spots that are created by these so-called acoustic oscillations in the early
universe are limited by how far a given dollop of energy and matter can
stretch or shrink in 380,000 years, which itself is limited by the speed of
light.
CHAPTER 7
INFORMATION (1995–1997)
There’s a photograph from Hawking’s wedding day in the early fall of 1995
in which the physicist and his new bride touch foreheads together. He sits in
his wheelchair in a dapper gray suit, half-smiling at Elaine, who smiles back
at him. The physicist’s right hand sits in his lap, cradling the clicker which he
uses to control his computer, and his wrist seems impossibly thin. The sleeve
of his suit looks almost empty. His left hand is on the armrest of his
wheelchair, cocked at an uncomfortable-looking angle, seemingly an attempt
to touch his bride. But Hawking had almost no control over his arms, and the
loving contact seems posed. His atrophied jaw and neck muscles can make it
hard to tell the difference between a smile and a grimace. The only reliable
hint at what is going on in Hawking’s head comes from his eyes, which gaze
at Elaine in a way that might indicate love, or anticipation, or even something
else entirely.
Hawking jealously guarded his innermost thoughts, seldom letting
information about his private life leak out to a public hungry for scandal.
Only those who knew and loved him could readily interpret his emotions and
gauge his well-being. Jane, of course, seemed of a mind that the physicist had
been manipulated by Elaine. (“He is in the grip of forces that he can’t control
and which broke up our home,” she insisted.) However, Jane had ceased
being his confidante half a decade prior—and, due in part to the acrimony of
their divorce, was not a reliable narrator of Stephen’s needs or desires.1
Stephen had announced his intent to move out with Elaine in late 1989,
but it was some time before he could get his affairs in order and find a place
to stay. As was to be expected for a feuding couple (actually, two couples)
living under the same roof, it got ugly—at one point, Jane writes, she put
Stephen’s suitcase outside the locked door in hopes that he would finally
leave; the result was a brick hurled through a window. Finally, in February
1990, Stephen and Elaine moved out, but the troubles didn’t end there. Tax
collectors pursued both Jane and Stephen regarding profits from A Brief
History of Time, and, as Jane put it, he was forced to pay a “monumental
fine” after “having first protested that he had already paid enough tax to build
a small hospital.” (“I remember a story where he went to give some talks in
Japan because it was tax time,” former student Ray Laflamme recalls. “I
don’t know how true it was, but I remember the students having some
rumors.”)2
Making matters worse, Jane writes, Stephen “did not instigate divorce
proceedings for a long time,” leading to a five-year delay before the divorce
became final in May 1995.i Jane was further hampered in her ability to move
on with her life by an “unjust and whimsical quirk of English law”: the
moment she remarried, she would lose the ability to claim a financial
settlement from Stephen.3 ii
However, had Jane consented to the divorce, Stephen would have been
able to initiate the divorce after having lived apart from Jane for two years.
Without that consent, English law required Stephen to wait until the fifth year
post-separation before filing.iii The timing of the divorce so soon after the
five-year rule allows, coupled with the speed of Stephen’s remarriage
afterward, suggests that Stephen was just as eager to get on with his life as
Jane was, if not more so.4
Jane had been closed out of Stephen’s life and had almost no information
about the physicist’s state of mind. In fact, she only found out about the
upcoming nuptials after the press did in July 1995. “I think he has been very
ill-advised in what he is doing, if he is planning that,” she told the papers
when they asked her for comment. “His present relationship is nothing to do
with me.” As she later wrote, “With his marriage, Stephen effectively
slammed the door on our remaining lines of communication.… I had no
choice but to reconcile myself to the end of an era.” Neither Jane nor their
three children attended the physicist’s wedding.5
The only person who could truly speak for Stephen was Stephen. Yet he
was all but silent, even though a number of his friends and former students
worried for his happiness and even his safety after his new marriage.
Whatever information there was was safely locked inside his head.6
Stephen Hawking’s work inhabited the very core of physics—and had even
raised the hope of generating a theory of everything, a single mathematical
framework that could reconcile the seemingly irreconcilable realms of the
quantum and the relativistic. His analyses of black holes required elements
from both of these incompatible worlds, and he had been able to find
profound new truths—that black holes radiated—by looking at the region
where those two theories clashed.
The most elegant expression of that clash was Hawking’s black-hole
information paradox: as Hawking interpreted it, the decay of black holes
pointed directly to a flaw in quantum theory, a place where the existing
framework in our understanding of the quantum realm needed to be patched.
This was the very sort of region that could give birth to a theory of
everything.
By the mid-1990s, however, physicists had trained new and powerful
mathematical tools on the black-hole information paradox and were
generating increasingly convincing evidence that Hawking’s interpretation
was wrong. The framework of quantum theory was withstanding his assault,
and the answer to his paradox was turning out to be considerably more subtle
than he had expected. It would be a number of years yet before he would
concede this point—and his famous wager—but the mid-1990s marked the
time that Hawking ceased being the leading expert on the paradox he himself
had formulated. A new and younger generation of physicists was generating
information about black holes that had evaded Hawking’s powerful intuition.
Hawking the symbol was transcendent, his place in the firmament all but
assured, even as Hawking the physicist was engaged in an intellectual
struggle—a battle that would bear directly on his scientific legacy.
Even if Hawking had been a native string theorist, the pace of his
communication gravely limited how much guidance he could offer his
students. As he gradually lost control over the last muscles in his hand, using
his computer by clicking the little pressure switch became more and more
difficult. By the late 1990s and early 2000s, he was slowly losing the ability
to talk even with the assistance of his computer. Galfard would attempt to
speed up the pace by looking at Hawking’s computer screen and attempting to
complete the sentence, using the expressions on Hawking’s face to guide
whether he was on the right track.33 xvi
Yet Hawking was confident that he and his students could point out flaws
in the string-theoretic models of black holes. Somewhere, he was sure, the
other physicists had made a mistake. Either their calculations or their
assumptions were wrong. Information couldn’t be preserved by black holes
—it had to be lost. He would stake his reputation on it. Publicly.
Hawking was famous for his wagers. In early 1997, when he made his
regular pilgrimage to Caltech in Pasadena, a wager was on the agenda. Two,
actually.
The first bet didn’t really involve information at all, but it did involve
black holes, or, more precisely, singularities, which reside at the heart of
every black hole.
The equations of general relativity assume that we all live in a nice,
smooth manifold of spacetime; it curves and ripples, but is never
discontinuous or pointy. It never misbehaves. Yet a singularity is a region in
spacetime where that assumption no longer holds—and that tends to break the
equations. The general-relativistic picture of black holes has a singularity at
the center—the curvature caused by gravity is so great that it effectively
punches a hole in the nice smooth surface of spacetime. But—and this is a
very interesting but—we can never see the laws of relativity break down.
Because the black hole has an event horizon around it—a surface beyond
which no information can escape, not even if it travels at light speed—no
information about the singularity leaks out. It’s as if the event horizon were
the boundary of a totally separate universe from the one we live in; nothing
from inside the horizon can ever enter our universe at all.
But are all singularities similarly shielded from prying eyes? Are all
places where general relativity’s equations break down hidden behind an
event horizon? In the late 1960s, Roger Penrose suggested that the answer
was yes—that singularities could never be “naked,” or directly visible.
Hawking, like Penrose, believed that there was some form of “cosmic
censorship” that made it impossible to observe a singularity directly. But
others weren’t so sure—like Caltech’s Kip Thorne and John Preskill. So in
1991, during one of Hawking’s visits, the three drew up a bet:
The wager was sealed with the two physicists’ signatures and one’s
thumbprint.
In the intervening time, however, gravitational-wave theorists had come
up with a weirdo scenario—one that wouldn’t occur in a real universe—
where a bunch of perfectly symmetrical gravitational waves collided at a
point in spacetime. The sloshing of space and time would be so violent at
that central point that if the energy were just right, it would create a
singularity without generating a black hole. Hawking was forced to concede
the bet—but he felt he had lost on a technicality, rather than a real
counterexample to the cosmic censorship principle. Indeed, the fact that the
energies had to be just so, and set up in just the right way to create even an
infinitesimal singularity, convinced Hawking more firmly than ever that he
was fundamentally right. But honor dictated that he concede. So he did.
At a public lecture, Hawking admitted defeat—but with a twist. Preskill
was giving the lecture, and Hawking, who was in town, introduced him to the
audience. “He said that I was the all-American boy because I like to drink
Diet Coke and like baseball,” Preskill remembers, “but then he wound up
saying that he was conceding the bet to Kip and me and he had T-shirts for us
to put on.”35
Thorne later wrote:
It’s not every day that Stephen gets proved wrong! With his
concession, Stephen gave each of us the promised article of clothing: a
T-shirt with his concessionary message. Sadly, I must tell you that
Stephen’s message… was not entirely gracious! He placed on the T-
shirt a scantily clad woman. (My wife and Stephen’s were aghast at
this, but Stephen has never been politically correct.…) [T]he woman’s
towel says “Nature abhors a naked singularity.” Stephen conceded, but
he asserted that Nature abhors that which he concedes Nature can
do.36
Two signatures and a thumbprint, and the wager was on. And it was clear
that Hawking thought he would win. Some of his students were busy learning
M-theory and preparing to fight back against the string-theory assault.
Unfortunately, they were about to get outflanked.
The second superstring revolution wasn’t over yet, and hard-core string
theorists like Ed Witten, Andy Strominger, and Juan Maldacena were filled
with enthusiasm as they produced important result after important result. Just
months after Hawking affixed his thumbprint to the bet, Maldacena, then only
twenty-nine years old, found another duality—a place where two seemingly
different mathematical models were actually equivalent—that landed the
death blow on the idea of black holes destroying information.
As with any duality, there are two pieces, two separate mathematical
objects, that wind up being equivalent in some fashion. In Maldacena’s, the
first piece is a well-studied form of spacetime known as anti–de Sitter space,
or AdS. AdS is a strangely shaped kind of spacetime that doesn’t look very
much like the spacetime of our universe, but obeys all the physical rules of
general relativity nonetheless. This hypothetical manifold serves as a
mathematical toy; physicists can use it to play with the equations of general
relativity (and string theory) in novel ways. For example, one can construct
black holes in an AdS space, and because of the peculiarities of the
manifold’s shape, a large black hole never evaporates as it would in a more
natural version of spacetime, whereas a small black hole does.
The second piece of Maldacena’s duality is conformal field theory, or
CFT. CFTs are mathematical frameworks that describe the behavior of
particles without mass. And, as it turns out, we’ve already encountered
something that’s almost a CFT: the Standard Model. The Standard Model
doesn’t quite meet the definition of a CFT (many particles in the Standard
Model have mass), but the underlying mathematical structures have a lot in
common. Think of a CFT as a supercharged Standard Model that describes
every particle’s interaction in a universe without mass—and, consequently,
without gravity.
Maldacena’s 1997 paper suggested that these two very different-seeming
mathematical structures were, in fact, dual to each other. Examining strings in
an AdS universe with spacetime curvature and black holes and other effects
of gravitational theory was essentially equivalent to studying particles in a
CFT where gravity doesn’t exist. Thus, according to Maldacena, one could
build a black hole in an AdS space and then bop over to the tools of the
corresponding CFT to analyze it.
This is a very abstract and counterintuitive concept. By studying the
behavior of particles moving about in a universe without gravity, one can
understand how black holes behave in the warped spacetime of a universe
where gravity is of prime importance. That is, the AdS/CFT duality allows
one to understand black holes without having to deal with the troublesome
mathematics of singularities in gravitation. And within a couple of months of
Maldacena’s paper, Ed Witten published a follow-up paper where he did just
that.xv
Looking at a swarm of particles in the CFT space is exactly equivalent to
observing a black hole in the AdS space—and just as particles have no
special ability to wipe out information in the CFT space, black holes
shouldn’t be able to destroy information in the AdS space. You could build a
tiny black hole in AdS, dump information into it, let it evaporate, and, voila!
The information would still be there in the CFT side of things, so it had to
still be there on the AdS side.
The AdS/CFT correspondence doesn’t directly say anything about real
black holes in our non-mathematical-toy universe. And it wasn’t even a
proven fact, but a conjecture. Yet the Maldacena paper rocked the world of
physics, as did the Witten analysis. For those battling over information loss
in black holes, the burden of proof had shifted—there were convincing
arguments that black holes couldn’t erase information in various toy
environments, and there was no reason to believe that a real black hole
would behave any differently. If, for some reason, our universe behaved
differently from all the other toy universes, that required explanation.
To some, the argument was over. “The moment I saw the Witten paper, I
knew the Black Hole War was finished,” physicist Leonard Susskind wrote.
“Whatever else Witten and Maldacena had done, they had proved beyond a
shadow of a doubt that information would never be lost behind a black hole
horizon.” Even Hawking was shaken, though he continued to insist that black
holes destroyed information. He continued pressing his students to brush up
on M-theory and the latest thinking about AdS/CFT, and to figure out
convincing counterarguments that could preserve the hope of information
loss. Yet they couldn’t help the master out of his bind.39
“I think it became clear to him that he was wrong, and he didn’t want to
go down on the wrong side of history on this very important question,”
Strominger says. And so, in 2004, Hawking finally convinced himself—using
his own mathematical constructs—that information could not be lost in a
black hole. He realized that he was wrong, but in true Hawking form, he only
came to that conclusion on his own terms.40
This meant he had to settle the bet. Preskill asked for a copy of Total
Baseball: The Ultimate Baseball Encyclopedia, but in baseball-free Britain,
that was a tall order. “At one point, we were having trouble getting it,” one
of Hawking’s assistants told a newspaper, “and he tried to persuade John
Preskill to take an encyclopedia of cricket, which of course we could find in
England, but John Preskill is a baseball fan, being an American, so that
wasn’t good enough.”41
Hawking took to the stage in Dublin to take his medicine and concede the
wager. Preskill had, indeed, bet wisely; he had (probably) been correct about
the answer to one of the most vexing physics questions of the late twentieth
century. And Hawking was (probably) wrong on a subject that he had spent
three decades of his life working on. Hawking’s late “solution” to the
information-loss paradox presented at Dublin convinced almost nobody—it
was almost as if Hawking simply couldn’t stand the idea of not answering the
question himself.
“I think one has to put that in the context of a physicist wanting to be at the
top of the game, wanting publicity,” says Taylor. “For Stephen, being well
appreciated, scientifically, was important to him, not just for his sort of
science but also for his health—to give him motivation to carry on—and also
in terms of his sponsors.”42
Not getting the right answer to the black-hole information paradox should
have been no source of shame. After all, Hawking was assured of his share
of glory merely by asking such a profound and important question in the first
place.
By this time in his life, Stephen Hawking was used to losing bets, but this
one was different. The bafflement and bemusement that surrounded the
concession of the black-hole-information-paradox wager showed more
clearly than ever before that Hawking was no longer at the intellectual center
of physics. Even though he himself had formulated the paradox, Hawking’s
ruminations on the subject were not cutting edge.
Dublin was a graphic demonstration: there was no escaping the
conclusion that Hawking the scientist was no longer of the first rank. Those
days lay behind him. To see his once-great scientific mind in its full glory,
one must turn back the clock even further.
Footnotes
i The petitioner on the divorce decree was technically Jane, not Stephen.
ii Stephen, who apparently owned the intellectual property that formed most
of the couple’s assets, was free to marry his nurse; Jane, on the other hand,
had to wait to marry her lover until after she officially had claim to some of
that money. According to Jane, the settlement took eighteen months; she and
Jonathan Hellyer Jones married in 1997.
iii There are other grounds for divorce, but Stephen would not have been
able to avail himself of them. He couldn’t have claimed desertion or abuse.
Even if it were perfectly clear that Jane’s relationship with Jonathan was
adulterous, the law said that excuse was only valid within six months of
Stephen’s finding out—and Stephen had known about the relationship since
the late 1970s.
iv Quantum information is somewhat more complicated than classical
information in a number of ways. Because a quantum particle, unlike a
marble, can be in two (or more!) places at once, a quantum bit isn’t just a 1
or a 0 like a classical bit, but can be both at the same time. Also, a quantum
bit can’t be copied without destroying the original—so while it’s true we
could make a perfect copy of the electron-in-box system described here, it’s
only possible to do so if the original electron-in-box system is ruined. So it’s
not really a “copy” as much as it is a “transfer” of quantum information.
v Actually, they explode in an extremely violent eruption.
vi Einstein was such a celebrity that he didn’t even have to show up in
person to draw Beatles-sized audiences; three years earlier, a crowd of
4,500 eager viewers trying to see a film about Einstein and his theory of
relativity stormed the American Museum of Natural History in New York.
Wags dubbed it the first “science riot.”
vii When he eventually agreed to do the lecture, Hawking apparently donated
all the proceeds to a charity for ALS.
viii At core, these frameworks describe the symmetries of abstract objects in
space. A bigger object looks more “complex” because it allows for (and
generally requires!) more particles and ways they can interact. But it’s
“simpler” in the sense that you only need one object to fully explain
everything rather than having to combine several different objects or theories
together to examine the same phenomena.
ix In 1984, physicists John Schwarz and Michael Green figured out that with
two particular types of symmetry, SO(32) and E8×E8, all the ugly elements in
the theory that might give rise to anomalies—breakdowns in the theory—
miraculously canceled. This gave theorists the green light to try to use string
theories as a foundation for a theory of everything.
x What the “M” stood for was never specified. Two possibilities are
“matrix” or “membrane,” but it may be something else entirely.
xi In this case, “simplified” means “five-dimensional and carrying lots of
electric charge.” This made the calculations easier.
xii Technically, they calculated the black hole’s entropy, a concept intimately
related to the concepts of temperature and information in ways that will be
made clearer in Chapter 13.
xiii Intel’s press release put even more effusive advertising copy in the
professor’s (computerized) mouth. It read: “‘Intel’s newest Pentium
processor technology keeps me connected to the world,’ said Stephen
Hawking, Lucasian Professor of Mathematics at the University of Cambridge.
‘I have immediate access to the Internet and email wherever I am. I must be
one of the most connected people in the world, and I can truly say, I’m Intel
inside.’” Intel, “Professor Stephen Hawking Stays Connected to the World
Through the Latest Intel Technology,” press release, March 20, 1997,
www.intel.com/pressroom/archive/releases/1997/CN032097.htm.
xvi The hand-operated clicker would soon be replaced with a sensor that
detected the twitching of Hawking’s facial muscles. And Hawking could
communicate to some extent, especially with people who knew him well,
with his facial expressions; this became increasingly important as he grew
older and the speed at which he could use his computer slowed.
xv And even more. The Witten paper emphasized that the CFT theory lived
on the boundary of the AdS space—think of CFT as a box and the AdS as
living on the inside of the box. AdS/CFT showed that all the information that
resides in the inside of the box is exactly the same as the information that is
inscribed on the box itself. This is a profound concept known as holography.
CHAPTER 8
IMAGES (1990–1995)
The secret was out. “Wheelchair Physicist in Love Tangle with Nurse,”
blared the Daily Mail. Though Stephen had moved out of the household in
February 1990, it wasn’t until July that the tabloids got wind of the
separation, thanks to an accident. Stephen’s wheelchair was knocked over by
a car—a surprisingly common occurrence for the physicist, who had a
reputation for fearlessness when driving his wheelchair—and suffered a
broken shoulder. As the press investigated the accident, they quickly realized
that Stephen had changed addresses. Reporters, with their noses finely tuned
to any scent of scandal, figured out that he had separated from Jane.1
Some newspapers looked for a cause. “The split partly came about
through religious differences,” an (anonymous) former student told a different
paper. “She is a committed Christian and as he increasingly became
interested in scientific rather than religious explanations, it became more
difficult for them. They are still the best of friends but in those conditions you
can’t expect people to have a normal married life.” Others looked for a
victim. The Daily Mail chose David Mason, Elaine’s husband and the father
of her two children. David also happened to be the engineer who had
adapted Stephen’s computer speech system so it would fit on his wheelchair.
“The whole situation is ludicrous and bizarre,” he told the Mail.2 i
The attention was unwelcome, and not just because of the embarrassment
involved. It effectively notified their landlord, Cambridge University, that
Stephen had moved out of the couple’s house. “Once the separation had
entered the public domain, the College lost no time in sending the Bursar
across to enquire when we were going to move,” writes Jane. “He was quite
explicit: The College felt itself under no obligation to house the family if
Stephen, with whom the College had signed the agreement, was no longer
living there.” Eventually, the university agreed to give Jane; her eleven-year-
old son, Tim; her seventeen-year-old daughter, Lucy; and her partner,
Jonathan, a year to find other accommodations.3
Stephen and Jane hadn’t expected to reach their twenty-fifth wedding
anniversary; it was almost miraculous that Stephen had survived to see their
fifth. That didn’t make it any less painful when their bond had finally
sundered. Jane still clung to the hope that they could preserve the marriage—
a hope that Stephen fed when he told the press there remained “a chance of
reconciliation.”4
Even so, Jane would have to move.
In 1990, Stephen Hawking was trying to adjust to his newfound fame. He had
been well known within the scientific community for decades, but the
professor’s sudden popularity was something entirely different—he had
become the world’s most recognized scientist. More, he had become a
symbol to the public, a transcendent mind in a withered body. And as much
as Hawking wanted to be known for the former, the latter was just as much a
part of his public persona. And as a celebrity, he needed to cultivate and
maintain that persona even when it didn’t capture the complexity of the
human underneath.
Luckily, Hawking had a genius, a natural talent, for nurturing his
newfound celebrity. His wicked self-deprecating wit, coupled with an
overweening certainty about his understanding of the natural world, made
him into the perfect archetype of the scientific genius. He succeeded, even
though for the first time antagonists were quite publicly trying to knock him
down a peg or two.
By 1990, the year Stephen moved in with Elaine, A Brief History of Time
had been on the best-seller lists for almost two years and was still selling
strong. Not only had the book made the physicist a publishing phenom—and
was beginning to earn him a tremendous amount of money in royalties—it
had made him an international celebrity. Though fame was new to him, the
limelight didn’t feel uncomfortable.
“People are either drawn to it or they aren’t,” says Peter Guzzardi, the
editor of Brief History. “What is that intangible something… that unqualified
impulse to step under the big Klieg lights and do your thing? I think Stephen
definitely had that. He loved the media spotlight, and it just—he came alive.
He’s one of those people that just came alive when he was the center of
attention. As well he might, you know; it must have been tough, just being
Stephen.”5
Hawking’s celebrity, or, more precisely, the wealth that accompanied it,
ameliorated the physicist’s difficulties finding and affording sufficient care to
keep him alive and able to work. Since his tracheostomy operation in 1985,
he needed 24-hour care, 365 days a year, and, according to Jane, “only a tiny
fraction of this expense would be borne by the National Health Service.”
Typically, ALS sufferers would require home care, institutionalization, or
hospice care only for a relatively short time; the system was simply not set
up to deal with the singular presentation of the professor’s affliction. His
longevity—a surprise to everyone, including himself—turned what was
usually a short-term situation into something that required a long-term
solution.6
In the late 1980s, the Hawkings had received help from the John D. and
Catherine T. MacArthur Foundation and other charitable sources—
administered through Cambridge University—to help pay for the nurses. Jane
estimated that the £36,000 per year they received—worth about $170,000 in
today’s US dollars—“just covered the bills.”ii As Stephen’s condition
worsened, the cost steadily climbed. However, now that Hawking’s book
was a success (and he was married to his chief nurse), the financial pressure
that he and his family had been under was greatly reduced. And he was able
to start turning his attention to other people in worse straits than himself.7
Hawking had particular empathy for children who suffered physical
disabilities; unlike them, he had had a relatively normal, able-bodied youth,
only to be afflicted in early adulthood. “It is very important that disabled
children should be helped to blend in with others of the same age,” he told an
audience in 1990. “How can one feel a member of the human race, if one is
set apart from an early age? It is a form of apartheid.” Hawking certainly
chafed at being set apart from his peers because of his disability instead of
his brain; he saw nothing heroic in his struggle against his disease. “I find it a
bit embarrassing in that people think I have great courage,” Hawking told a
reporter in mid-1990. “But it is not as if I had a choice and deliberately
chose a difficult path.”8
Hawking typically made light of his disability, and he denied that it
hindered him in his work. “I was lucky to have chosen to work in theoretical
physics because that was one of the few areas in which my condition would
not be a serious handicap,” he said at an ALS conference in 1987. Not only
didn’t his disability hinder him, he would tell his audience, but it actually
helped him. “In fact, in some ways, I guess it had been an asset: I haven’t had
to lecture or teach undergraduates, and I haven’t had to sit on tedious and
time-consuming committees,” he wrote. “So I have been able to devote
myself completely to research.” And that’s on top of compensation for not
being able to write: his enhanced visual sense.iii “If you can’t write it, you
become better at holding diagrams in your brain,” cosmologist Alan Guth
told sociologist Hélène Mialet in 2005. It was part of the mystique; Hawking
had transcended his body. What would be a crushing blow to mere mortals
had not only failed to defeat Hawking, it had strengthened him, given him
power.9
In public, Hawking’s stoicism almost never wavered. Almost. In 1988, a
journalist asked the scientist whether, given the option, he would want to
“[regain] the power to walk and feed himself at the expense of being
intellectually mediocre.” Hawking’s surprising reply: “Yes.” Then Hawking
backtracked. “I don’t want to be anyone else. People should be who they
are.”10
After Hawking lost the use of his hands, he had to do calculations entirely in
his mind. And his was a geometric mind.
In mathematics and physics, there are often multiple ways to solve the
same problem. Actually, “solve” isn’t quite the right word. Mathematicians
grasp for something deeper than a solution: a profound understanding, a
weaving of the answer into one’s intuition, an incorporation of the essence of
the problem directly into one’s brain so that it becomes a part of you. You’ll
often hear mathematical types try to describe this feeling with the term grok
—a transcendent form of understanding described in Robert A. Heinlein’s
science fiction novel Stranger in a Strange Land.
Pure mathematicians or physicists don’t seek solutions as much as they
seek this grokking of something beyond themselves, an extension of their own
brains into new domains—domains that become brand-new mental
playgrounds, new sources of things to explore.
Different mathematicians have different means of grokking a problem.
Some might seek understanding in patterns and symbols, and the relationships
among them and the rules for manipulating them: this is an algebraic
viewpoint. It is a world of formulas and numbers and functions and
operations. Others might try to understand problems by drawing mental
pictures: seeing shapes in space and how those shapes move and interact
with and relate to each other. This is the geometric viewpoint. For example,
an algebraic thinker might envision a square number as a number that can be
expressed as n × n, an integer multiplied by itself. A geometric thinker might
envision a square number as dots that can be perfectly arranged into a square
pattern; only numbers that can be so arranged are square numbers (hence the
name).
These two viewpoints aren’t mutually exclusive—they bleed into and
reinforce each other. And even if a mathematician tends to gravitate toward
one viewpoint or the other when trying to grok a new area, he or she can use
algebraic techniques to survey the landscape and build up a deep geometric
intuition later, or vice versa. That’s one of the beautiful things about
mathematics: one can approach the same idea in many different ways, each of
which might give a different perspective and understanding.
To understand how a body moves through general relativistic spacetime,
an algebraic thinker would examine the equations of general relativity. By
manipulating the symbols in the equations—representing mathematical
objects called tensors that encode the warp and weft of spacetime—he could
figure out the object’s path. A geometric thinker who has properly trained her
mind’s eye might simply be able to “see” the object moving and conclude
how it rolls around on the abstract four-dimensional surface. These are both
valid ways to try to understand general relativity. Sometimes (but not
always), geometric thinking is more intuitive and algebraic reasoning more
precise, but both approaches are powerful—and natural—ways to try to
understand the cosmos.
The rules of spacetime, in particular, are naturally suited to a geometric
thinker. After all, the algebraic formalism of general relativity, the equations
describing how gravity works, describes curvature—a geometric concept. It
might be a funny sort of curvature in four dimensions (or even more
dimensions in toy models of the universe!), but it is still something
fundamentally geometric, something that could be understood reasonably
well by drawing mental pictures. If you think in the right way, you can “see”
the rules of Einstein’s spacetime. And a diagram of spacetime is a good
example: If you can imagine your path through spacetime as a little dot
moving upward (through time) and to the right (through space) on the
diagram, the dictum that you can’t move faster than the speed of light is
exactly equivalent to saying that you can’t take a path on the diagram steeper
than 45 degrees from the vertical. Thinking geometrically, you can
immediately see that there are regions in spacetime that you’re not allowed to
reach by virtue of the light-speed limit, and you can immediately divide the
universe into two: the parts you can reach, and the parts you can’t. If you
grasp these kinds of geometric rules, you can train your mind to grok
Einstein’s theory. You might even be able to make discoveries.
Even before ALS began to take its toll, Hawking seemed to be attracted to
geometric modes of thinking, which may be part of the reason general
relativity appealed to him in the first place. As his disability became more
profound, he shifted radically to the geometric side of the spectrum, avoiding
algebraic symbol-manipulation whenever he could. As he said in 1998, he
typically tried to “avoid problems with a lot of equations or translate them
into problems in geometry. I can then picture them in my mind.” Without the
ability to write things on paper, without an easy way to keep track of
complex manipulations of symbols, he was largely foreclosed from making
new discoveries using heavy-duty algebraic reasoning. At best, he could
show his students the instruments of calculation and let them do the work. But
when it came to geometric reasoning, he could more than hold his own. And
as he developed his mind’s eye to handle the geometric intricacies of general
relativity, he was able to hone his intuition—and be led by it. As his friend
Kip Thorne wrote, “Hawking is a bold thinker. He is far more willing than
most physicists to take off in radical new directions, if those directions
‘smell’ right.” Or, as Thorne knew very well, to throw cold water on an idea
if it smelled wrong.11
In the summer of 1985, the astronomer Carl Sagan sent Thorne a draft of
his novel Contact—in which Earth travelers had to use an alien craft to
travel almost instantly to the center of the Milky Way and back—and asked
for Thorne’s help coming up with a way to do that. The light-speed limit
would seem to prevent that; there was no way the travelers could make a
round trip to the center of the galaxy (about thirty thousand light-years away)
in fewer than sixty thousand years (as measured by a clock on Earth). It was
equivalent to asking to find a path into the “forbidden” region of a spacetime
diagram; there was no way to do it without, at some point, having a line
extend more than 45 degrees from the vertical. However, that assumes that
spacetime is a nice, smooth, unbroken sheet. What if the sheet had a hole in
it? A black hole with its singularity?12
“I have my crew going through a black hole, and I know you’re not going
to like that,” Sagan told Thorne. “Can you help me rework it?” Inspired by
Sagan’s question, Thorne and his graduate students started calculating.iv And
Thorne quickly concluded that a black hole couldn’t be used as a means of
transport. But there was another possibility: instead of a singularity, a point
of infinite curvature, what about a smooth, non-singular tunnel? A wormhole?
13
“I rely on intuition a great deal,” Hawking told a BBC radio host. “I try to
guess a result, but then I have to prove it. And at this stage, I quite often find
that what I had thought of is not true or that something else is the case that I
had never thought of.”20
In the early 1990s, A Brief History of Time was still setting sales records,
and the press couldn’t get enough of the new scientific superstar. Everyone
was trying to get inside Hawking’s head, to try to figure out how his mind
worked. Not since Einstein had the public been treated to such an archetype
of pure intellect, a superlative mind that appeared to have such a tenuous
connection with the worldly plane of everyday existence. Hawking had been
appreciated—and highly honored—within his field for years. But with the
publication of A Brief History of Time, he had become a household name. A
celebrity. And, to some, a hero. The demands on his time were unceasing:
there were talks, interviews, invitations, and honors.
Other schools, including his beloved Caltech, tried to lure him away from
Cambridge. “Trying to pry him away from the Lucasian Professorship was
certainly a long shot, but Kip and I went on a secret mission in 1991,” says
John Preskill. They offered him a newly minted professorship, the Richard P.
Feynman Chair, but in the end Hawking decided to stay in England. “The
outcome of the discussions was that we agreed he could come and visit every
year. These visits were quite expensive, because he would bring a whole
entourage—his medical team, and he would bring students—but we got a
foundation, the Sherman Fairchild Foundation, to pay all those expenses and
to agree to do so on an ongoing basis.” Hawking was in such demand that
people would break the bank to get a little piece of him.21
Stephen loved the attention—as did Elaine, who almost literally bounced
off the walls with excitement at her soon-to-be husband’s appearances. (She
would reportedly do cartwheels in odd places, such as at a reception where
Stephen received a degree at Harvard University, or on a film soundstage
where Hawking was being interviewed for a movie.) But the professor
steadfastly denied that the worldly comforts of fame and fortune had changed
him—if anything, he said, his newfound celebrity was an inconvenience and
a distraction.22
“It has not made much difference,” Hawking told an interviewer for
Playboy in 1990. “Even before the book, a certain number of people, mainly
Americans, would come up to me in the street, but it has made that sort of
encounter more frequent. And other things like interviews and public lectures
have taken up the limited time I have to do research. However, I’m now
cutting down on such things and getting back to research.”23
Even in the early 1990s, when Hawking’s international fame was still
new, his interviews had the practiced polish of a man who was comfortable
with media attention. Part of this was because journalists simply couldn’t get
unfiltered, spontaneous thoughts out of the physicist. At best, Hawking’s
interviewers couldn’t get more than a question or two in in real time, and by
virtue of the labor and time that Hawking had to expend in crafting those
responses, they were more fussed over and deliberate than most off-the-cuff
statements by other celebrities. Most of the time, even that wasn’t an option;
because the physicist communicated so slowly, the interviewers had to
submit their questions in advance. Hawking and his assistants had the luxury
of time to craft the perfect answer to each question—a response that would
burnish the image of the physicist that he himself was building.
The image the public had fallen in love with was not an easy one to create
or maintain. Hawking had to be humble at the same time that he was being
portrayed as the intellectual heir of Newton and Galileo and Einstein. He
couldn’t be defined by his disability even though the way he so graciously
coped with ALS was a major reason for his immense popularity. And he had
to be a success—a wealthy celebrity—while playing the role of physics
ascetic, eschewing worldly things in favor of cosmic knowledge.
In the Playboy interview, which appeared in April 1990, Hawking
couldn’t seem to settle on whether his origin story should emphasize how he
coped with his disability (“I chose my field because I knew I had ALS.
Cosmology, unlike many other disciplines, does not require lecturing.”) or
not (“From the age of twelve, I had wanted to be a scientist. And cosmology
seemed to be the most fundamental science.”)v And even as he insisted to the
Playboy interviewer that the publication of Brief History had made almost
no difference in his life, the repercussions were already immense; after all,
he was in the brief interregnum between having left his wife of a quarter
century and the press finding out about it.24
Jane, for her part, was crafting a counternarrative distinct from Stephen’s.
Jane, too, was trying to walk a fine line with her image of her husband. In it,
she had recoiled from a Stephen who was arrogant, selfish, and spiritually
crippled; the single vision of his hyper-logical worldview had robbed him of
some of his humanity. Jane emphasized his brilliance—after all, his fame,
and, by reflection, hers, was founded on the quality of Stephen’s brain. At the
same time, Jane implied that Stephen was precisely the opposite of
transcendent. Within Stephen’s brain spun the motes of suns and worlds and
spaces, yet his logic confined him to the dust, ciphering things beyond his
ken.
And Jane knew it even if Stephen didn’t. Not long before the separation,
she presented this image to a sympathetic reporter. “In the beginning, I felt
Stephen was a scientist and he shouldn’t involve himself in areas that didn’t
really concern him. Now he is coming up with such astounding theories…
that can have a very disturbing effect on people… and he’s not competent,”
she said in 1988, shortly after Brief History first came out. “But I pronounce
my view that there are different ways of approaching it and his mathematical
way is only one way—and he just smiles.” After the divorce settlement in
1995, she dialed up the volume. “A spiritual home as well as a dependable
family home is essential for every person born into this world,” she would
write in the first version of her memoir. “It is at our peril that we neglect
those deep-seated needs in favour of materialism, egoism, science or the
extremes of rationality.”25
It was a battle of well-worn tropes. Stephen’s tended to carry the day, for
they were much more compelling. And Stephen was a master at crafting his
image—and subtly helping others do it for him.
In 1992, Hawking was the guest on an unusually lengthy episode of the
BBC radio show Desert Island Discs. Sue Lawley, the host, would chat with
celebrities and talk about what music they would want to have if they were
stranded on a desert island. “In many ways, of course, Stephen, you are
already familiar with the isolation of a desert island, cut off from normal
physical life and deprived of any natural means of communication,” Lawley
said to him, straight off the bat. “How lonely is it for you?”26
“I don’t regard myself as cut off from normal life, and I don’t think people
around me would say I was,” he responded. “I don’t feel a disabled person
—just someone with certain malfunctions of my motor neurons, rather as if I
were color blind.”
As for Hawking’s choice of music, he couldn’t resist a selection from Die
Walküre, because Wagner “suited the dark and apocalyptic mood [he] was
in” when diagnosed with ALS. A Puccini aria, some Brahms, some Poulenc,
and a light Beatles song accounted for four more of his eight choices. His
remaining three were just as on-message as Wagner. A Beethoven string
quartet played by a character in a novel who knows he’s about to die. The
beginning of Mozart’s Requiem, the composer’s last, unfinished, work before
his death at the age of thirty-five. And Edith Piaf’s “Je ne regrette rien.”
“That just about sums up my life,” Hawking said. It was a powerful message.
So powerful was it that Hawking could make people choke up even with
television commercials. In 1993, Hawking was the star of a minute-and-a-
half-long television advertisement for British Telecom—the UK equivalent
of Ma Bell. In his robotic voice, Hawking begins, “For millions of years,
mankind lived just like the animals. Then something happened which
unleashed the power of our imagination. We learned to talk.” Hawking’s
wheelchair slowly drives under Mayan vaults and into a Greek amphitheater
as the physicist extols the virtues of communication. As a giant BT-branded
radio telescope fills the screen, Hawking concludes with a signature touch.
“With the technology at our disposal, the possibilities are unbounded. All we
need to do is make sure we keep talking.”27
“I saw an advert on the television in England, for a telephone company,
and [Hawking’s] voice was on this advertisement. And this advertisement
nearly made me weep,” David Gilmour, guitarist and vocalist for Pink Floyd,
told a radio interviewer a year later. “I’ve never had that with a television
advertisement before, or with a commercial on the television.… I just found
it so moving that I felt I had to try to do something with it.” So Gilmour took
Hawking’s voice from the BT advertisement and made it into a track on the
newest Pink Floyd album, The Division Bell.28
As Gilmour plainly saw, any performance that can make people teary-
eyed during a telephone commercial—that’s something unusual. That’s art.
At times, the line between science and art is hard to see. Theoretical
physicists, like artists, are often guided by a sense of aesthetics, a desire to
grasp something beautiful that feels just out of reach. A discovery, a paper, a
result, will often reveal but one aspect of a deeper truth that the physicist
feels, but cannot yet express fully. But there is, of course, a difference
between art and science: nature. Nature is the final arbiter of not just what is
true, but what is beautiful. The most aesthetically pleasing models wither and
rot if they are contradicted by experiment. Conversely, ideas considered ugly,
even revolting, come to be considered beautiful if they explain the way the
natural world works. Quantum mechanics, relativity, even atomic theory
were all rejected by eminent scientists who simply found the theories
distasteful. Over time, new experiments shape scientists’ ideas of how the
universe works, bending their aesthetic sense into alignment with natural law,
turning once-unthinkable ideas into paragons of beauty and elegance. Even
when that beauty is gathering dust.
By the late 1980s and early 1990s, the fields of general cosmology and
relativity were in something of a rut. There hadn’t been any real breakthrough
experiments or observations since the late 1960s and early 1970s. As a
consequence, both fields were ripe for experimental discovery; cosmologists
and gravitational physicists were hankering for some new observation that
would test their intuition. And prospects for something new in cosmology, at
least, were extremely high. The year 1990 saw the launch of the Hubble
telescope, the most capable orbiting telescope the world had yet seen.
Hubble would allow astronomers to look deep into the cosmos to see very
faint and distant objects, and to set better limits than ever before on the rate
of the expansion of the universe, as well as the conditions in the young
cosmos. And 1989 saw the launch of COBE, the Cosmic Background
Explorer. At long last, physicists would begin to see the hot and cold regions
in the cosmic microwave background that theorists had predicted, but that no
one had yet spotted.
General relativity was having an even tougher time. There was a bright
spot, however, with the 1991 launch of a NASA telescope designed to detect
gamma rays—light with even higher energy than X-rays. While gamma-ray
astronomy didn’t directly test theories of gravity, gamma rays, like X-rays,
might give glimpses of events so violent that they could only be the
handiwork of black holes. Indeed, a certain mysterious kind of gamma-ray
emission discovered in the 1970s seemed to hold the key to one of
Hawking’s early predictions, one that carried Hawking’s best chance of ever
winning a Nobel Prize.
The gamma-ray mystery first emerged in the early 1970s, thanks to a set of
secret satellites launched by the US Air Force. The idea was that if someone
detonated a nuclear bomb in the atmosphere or in outer space, these
instruments—pairs of bizarre-looking, black, twenty-sided things studded
with sensors—would detect the flash of gamma rays in the bomb’s fireball.
But unexpectedly, even though there were no nuclear explosions, the
satellites were spotting brief bursts of gamma rays from deep space. Such
bursts would have to be born in an extremely violent event—something like a
supernova in which a star collapses into a neutron star or a black hole—but
searches for supernovae that might be sources of the mysterious gamma-ray
bursts came up empty. If the bursts weren’t caused by the collapse of a star,
what other event could be responsible?
Stephen Hawking thought he had an answer. In 1976, along with one of his
students at the time, Don Page, Hawking argued that the solution to the
mystery might lie not in supernovae, not in the birth of black holes, but in
their death. For several years, he had been arguing that the universe was
populated with mini black holes—only as massive as a mountain or an
asteroid, rather than a star—created in the aftermath of the Big Bang. If they
existed, these primordial black holes should be ending their lives in massive
explosions. All around us, there should be mini black holes exploding,
releasing a blast of gamma-ray radiation every time one gives up the ghost.
Page and Hawking calculated the very rough properties that these gamma-
ray bursts should have if they in fact came from primordial mini black holes.
A definite observation of gamma-rays from a primordial black hole would be
a tremendous vindication of general relativity and quantum theory and would
give us important information about the early universe and strong interactions
at high energy that could not be obtained in any other way,” the pair wrote.
Unfortunately, the equipment of the day wasn’t good enough to make the
necessary measurements. The physicists realized that sort of observation
would have to wait until engineers could build a gamma-ray telescope with
high resolution, and fly it above the gamma-ray-absorbing atmosphere.29
In 1991, NASA launched the Compton Gamma Ray Observatory, an
enormous (and extremely expensive) orbiting gamma-ray satellite with
instruments that could map out where these mysterious gamma-ray bursts
were coming from and what their properties were. If the telescope had found
gamma-ray bursts that were the signature of mini black holes, it would have
very likely led to a ceremony in Stockholm for Hawking. “People have
searched for mini black holes of this mass, but have so far, not found any,”
Hawking told audiences years later. “This is a pity, because if they had, I
would have got a Nobel prize.”30
Unfortunately for Hawking, they hadn’t, and he didn’t. The Compton
satellite and its successors, coupled with LIGO, implied that some gamma-
ray bursts (the short ones) came from neutron stars slamming into each other
to create a black hole; others (the longer ones) do, indeed, seem to come
from distant, extremely powerful supernovae. Primordial black holes were
not the answer.
Hawking would begin to sour on the possibility of primordial black
holes. By 1993, along with a Cambridge colleague, John Stewart, Hawking
used a toy model of black holes—black holes in a two-dimensional universe
rather than our own four-dimensional one—to investigate what happened
when a black hole evaporated. In the model, one of two things happened,
either of which was bad news. The first was that it would create a naked
singularity, an open wound in the fabric of spacetime. Hawking was
absolutely convinced that this was impossible. (More on this shortly.) The
other possible outcome was what they dubbed a “thunderbolt”: a breakdown
in the equation that they interpreted as a violent burst of high-energy
particles. What’s more, Hawking and Stewart reasoned that the same sorts of
energetic explosions would happen in our four-dimensional universe. But,
they admitted—possibly with a tinge of sadness—these gouts of particles
couldn’t explain the mysterious gamma-ray bursts. “It would be tempting to
try to connect such events with the gamma ray bursts, but there is a problem
with the energies involved,” they wrote. Such events “would have to be
extremely mild and could not account for the observed gamma ray bursts. If
the universe does contain black holes that are reaching the end points of their
evaporation, it seems they will do it without much display.”31
Within a few years, Hawking seemed to be moving toward giving up on
the concept of primordial mini black holes altogether. One reason might be
the work that he did with one of his grad students, Raphael Bousso. In the
mid-1990s, Hawking tasked Bousso with calculating how many such mini
black holes would have been created immediately after the Big Bang.
“The project he actually gave me, he tried to explain in terms of some
kind of argument he was having with Roger Penrose,” Bousso explains. “He
basically wanted me to work on something that would show that the universe
doesn’t have to begin the way that Penrose says it should.… He basically
wanted me to calculate the probability that black holes would be created by
some kind of quantum process during a very early phase of the universe
called inflation, when space was expanding at an exponential rate.”32
“It was a new student, just starting with Stephen. And Stephen was
suggesting some problem for him to work on. And this problem was to show
that something that I had done was wrong. So the student had to do this,”
Penrose told sociologist Hélène Mialet in 1998. “Then I went to talk to
Stephen. And… I came out of the room, and I found the student waiting there,
rather nervously. And he came up to me, and he asked me, ‘What is it I’m
supposed to do?’ Because he… it was quite funny.” As Penrose later wrote,
“Being a student of [Hawking’s] was not easy.… Hawking might ask the
student to pursue some obscure route, the reason for which could seem
deeply mysterious. Clarification was not available, and the student would be
presented with what seemed indeed to be like the revelation of an oracle—
something whose truth was not to be questioned, but which if correctly
interpreted and developed would surely lead onwards to a profound truth.”33
Bousso soon got over his initial confusion, tackled the problem, and
began calculating the probability—which, in turn, would give an estimate of
how many primordial black holes would have been created. The answer was
pretty stark: none. Once the two plowed through the detailed calculations,
they realized that while there could possibly be lots of subatomic-sized black
holes forming, getting any black holes larger than that was exquisitely
difficult. The subatomic black holes would have evaporated away almost
instantly and wouldn’t have survived to the present day. There would
therefore be “no significant number” of primordial mini black holes left over
from the very early universe.34
The argument wasn’t perfect, and Bousso and Hawking made a number of
assumptions, particularly about the evolution of the universe in the first
moments after the Big Bang. Even without those assumptions, Bousso’s work
wouldn’t kill the idea of primordial black holes entirely. “There are different
kind of questions you could be interested in—Are you producing just some
kind of black hole in the universe, or specifically those kinds of black holes
that would finish evaporating now?” Bousso says. “If you just want some
kind of black hole, you have much more freedom.” To this day, scientists
look for signatures of primordial black holes (admittedly without really
expecting to find them).35
However, the idea of primordial mini black holes, which featured
prominently in 1988’s A Brief History of Time, didn’t make an appearance in
Hawking’s later works The Universe in a Nutshell or The Grand Design.
Even though Hawking was deeply attached to the idea—it was one of his
youthful theories and one that got him a lot of attention over the years—he
was willing to abandon his intuition in the face of greater evidence. Up to a
point.
Even if there were no primordial mini black holes from the first moments
after the Big Bang, that said nothing about mini black holes created on Earth.
As ludicrous as that sounds, it’s not out of the question: if engineers figure
out how to pour enough matter and energy into a small enough space, they
might be able to create a tiny black hole that appears and suddenly
evaporates away in a burst of energy. And pouring lots of matter and energy
into a tiny space is precisely what particle colliders like the Large Hadron
Collider at CERN are designed to do. “Some of the collisions might create
micro black holes,” Hawking said. “These would radiate particles in a
pattern that would be easy to recognize. So, I might get a Nobel prize after
all.”36
W hen a possible naked singularity appeared in one of his calculations,
Hawking knew in his heart that it couldn’t be; the alternative, no matter how
distasteful, had to be true. (The so-called thunderbolts were, indeed,
distasteful; they showed that the equations were failing.) This intuition went
back to his years as a student, and was directly inspired by the work of a
mathematician who was very influential on the young Hawking: Roger
Penrose.
Penrose was a decade older than Hawking, a young professor at the
University of London when Hawking was a student at Cambridge. Trained as
a pure mathematician, he got interested in cosmology thanks in part to
Cambridge physicist Dennis Sciama, and quickly made a splash by bringing
novel mathematical techniques to bear on problems in cosmology and general
relativity—including the problem of singularities. Along with Brandon
Carter, a student of Sciama’s at Cambridge, Penrose came up with a clever
method of visualizing what happens even in the seemingly unvisualizable
infinities that appear in the middle of a black hole.
The so-called Penrose (or Penrose-Carter) diagramvi begins with an
ordinary spacetime diagram like the ones in Chapter 4. Such a spacetime
diagram has infinities in it—you can imagine a path, say, that moves through
time without stopping, going into the infinite future, but it’s impossible
actually to plot such a path on the diagram. You’d need an infinite piece of
paper. But Penrose realized that you could use a mathematical trick in some
ways analogous to what painters use to portray unbounded distances in their
artwork: they create a vanishing point, a little dot in the middle of their
canvas, that is supposed to represent a place infinitely far away. Parallel
lines that all stretch off to infinity are all pinched together into that vanishing
point, and our brains automatically understand that the flat, finite image
represents something with infinite depth. Similarly, a Penrose diagram
pinches off all the parallel paths that stretch out through time, causing them to
converge at a point that represents the infinitely distant future.
Where the Penrose diagram gets more complicated than your average
perspective drawing is that there’s more than one vanishing point. Just as you
can imagine paths stretching to the infinite future, there are paths coming from
the infinite past—and the diagram helpfully includes a vanishing point for
that, too. But a spacetime diagram has infinities not just in time, but also in
space; you can travel in an infinite direction to the right or to the left, so there
are two more vanishing points for those two infinities. So, essentially, a
Penrose diagram is a spacetime diagram where the top, bottom, left, and right
are each pinched off into a vanishing-point infinity, making it look like a
diamond. And just as the corners of the diamond represent infinities, so, too,
do the lines that connect them; they can only be reached (if at all) by traveling
infinitely in one direction or another.
Images (1990—1995)
He told us that it was all right to study the evolution of the universe
after the big bang, but we should not inquire into the big bang itself
because that was the moment of Creation and therefore the work of
God. I was glad then that he did not know the subject of the talk I had
just given at the conference—the possibility that space-time was finite
but had no boundary, which means that it had no beginning, no moment
of Creation. I had no desire to share the fate of Galileo, with whom I
feel a strong sense of identity, partly because of the coincidence of
having been born exactly 300 years after his death!46
“He has a real chip on his shoulder,” Hawking told a reporter when asked
about Appleyard’s critiques. “I don’t know that I’ve seen him write
approvingly of anyone. I feel that he is a failed intellectual and so he has to
decry everyone else.”52
Hawking could easily dismiss attacks from people outside of his field;
after all, they didn’t have the training to fully understand his work. However,
Hawking had antagonists even within his own scientific community. John
Barrow, a scientific sibling—he, like Hawking, had had Dennis Sciama as
his PhD adviser—was a prominent voice in a brutal takedown of Hawking in
The Spectator. One passage read:
That wasn’t all. Barrow also implied that Hawking’s work, even the
“interesting” parts, would wind up being of no lasting value.
For a man struggling to transcend his mortality by seeking eternal
knowledge, there could have been no insult more keenly felt.
Footnotes
i The paper didn’t manage to get in touch with Jane, or catch wind of her
relationship with Jonathan Hellyer Jones. When reporters came calling,
“Jonathan, of whose existence they were unaware, managed to slip out the
back door.” The ruse was successful. “Jane, 48, is left to reflect alone,”
reported the Mail. Jane Hawking, Music to Move the Stars: A Life with
Stephen Hawking (London: Pan, 2000), 575; Emma Wilkins, “Wheelchair
Physicist in Love Tangle with Nurse,” Daily Mail, August 1, 1990.
ii The numbers have been converted to 2020 US dollars using late 1980s
exchange rates and forty years of US inflation.
iii However, an enhanced visual sense only extends so far. Hawking wrote,
“It is impossible to imagine a four-dimensional space. I personally find it
hard enough to visualize three-dimensional space!” Stephen Hawking, A
Brief History of Time (New York: Bantam Books, 1998), 24.
iv Sometimes involuntarily. Thorne put a question about wormholes on the
final exam of the fall 1985 session of his introductory course to general
relativity.
v The latter is closer to the truth; Hawking’s decision to study cosmology
couldn’t have been influenced by his disease because he wasn’t aware of his
illness at the time he chose his field. Hawking enrolled in Cambridge to
study cosmology and started his studies in October 1962; he was diagnosed
with ALS in January 1963. In Hawking’s words, “The doctors told me to go
back to Cambridge and carry on with the research I had just started in general
relativity and cosmology.” Given that the diagnosis meant that Hawking had a
two- or three-year life expectancy, he couldn’t have thought at the time that he
would survive long enough to advance far enough in his career to have to
lecture. Hawking, My Brief History, 47.
vi Technically, Carter diagrams, named after physicist Brandon Carter, are a
class of Penrose diagrams for spacetimes that are rotationally symmetric—
but even experts in the field don’t tend to make the distinction between the
two.
vii A quick example for the brave: Notice that the event horizon is at a 45-
degree angle, just like a light beam. This means that in some sense, the event
horizon is moving at the speed of light—even though it doesn’t get larger
when you observe it from the outside, from the point of view of someone
trying to escape from the black hole, the horizon is essentially receding at
light speed; it’s impossible to overtake it and escape.
viii With one glaring exception: the Big Bang.
ix Technically, Penrose presented a related conjecture about a set of
inequalities about mass and spacetime—if one found a counterexample to the
inequalities, it would automatically mean that the cosmic censorship
principle was wrong.
x There are plenty more that Hardy should have been aware of: Karl
Weierstrass, Joseph Fourier, Leonhard Euler, Pierre-Simon Laplace… and
that’s without even taking issue with his assumption of gender. Emmy
Noether was making major advances non-stop through her late forties and
early fifties and would no doubt have continued had she not died of cancer at
the age of fifty-three.
CHAPTER 9
FLASH (1987–1990)
The years 1987 through 1990 marked one of the greatest transitions in
Stephen Hawking’s life. At the start of this period, he was a well-respected,
if not terribly well-known, physicist who had tried his hand a few times—not
so successfully—at science popularization. At the end, he was one of the
world’s most recognizable international celebrities, a scientist superstar the
likes of which hadn’t been seen since Einstein.
A Brief History of Time, published in 1988, was the spark for his rise to
fame, even if it had little to do with his appeal to the public. There were
better-written, even better-selling popular science books out there, many of
which were written by first-rate scientists. But once Brief History hit the
shelves, the publishing world reeled. So did Hawking. His comfortable life
in Cambridge with Jane and the children, his career as a theoretical physicist
laboring in semi-obscurity, were both roiled by the sudden apparition of
fame and fortune and attention.
But for Stephen, the recognition couldn’t come soon enough. He had
worked for more than half a decade—nearly succumbing to his disease in the
middle—to make his dream of a physics best seller come to fruition. Now
nothing could stand between him and his desire.
There would be no more delays. It was five years since Hawking first
drafted the little one-hundred-page manuscript that would become the core of
A Brief History of Time. He had signed his contract with Bantam Books and
had begun discussing strategy with his editor, Peter Guzzardi, in mid-1984. A
year later, Hawking was in a coma in a hospital in Geneva. He would have to
learn how to communicate again, first without, and then with, a computer
embedded in his wheelchair. That work on the book began again in 1986 was
a testament to Hawking’s force of will. For there was a lot of work to be
done to turn that dense little draft into a book that people would want to read.
Or at the very least, buy.
“The material itself was not that promising, but the story of Hawking was
promising,” admits Guzzardi. “So the hope was we can take this material
which is kind of dry and uneven and sometimes written with all kinds of
assumptions about the reader already understanding a lot of physics and
astrophysics, and in some cases, written a bit like a middle-school level. So
it was very uneven.”2
The early draft of Hawking’s manuscript was uneven, perhaps, because
parts of it were derived from his prior lectures, most of which were intended
for physicists, or at least scientists. Hawking made an attempt to smooth over
the language and make it more accessible to the general audience, but he was
not always successful. For example, in the draft manuscript, he writes,
Other than eliminating some details and bumping others to a table, the
language is only cosmetically different, and is not the sort of material that
would appeal to a typical lay reader. Nor did the draft display any real
interest in the rich intellectual history of the field of cosmology. There were
few historical details, and none of the kind of humanistic information that
could put Hawking’s work in context, lighten the prose, and give the reader
some space to breathe in between difficult technical passages about
astrophysics and quantum mechanics.i
At one point, a frustrated Al Zuckerman suggested using a ghostwriter.
Guzzardi refused, but he did hire astronomer and science journalist John
Gribbin, who had known Hawking for a number of years (and later co-wrote
a biography of the physicist), to take an early crack at reshaping the
manuscript. “What I did with Peter—and Peter didn’t tell anybody else at the
time—was to rearrange it so it made more sense logically,” says Gribbin. “I
don’t want to say ‘disappointed,’ but I would have liked to have done more.
You know, if Peter had felt able, if Peter had actually asked him, Stephen
would have let me do it if he had known it was me, because we had a good
relationship. But Peter said that he promised that he wouldn’t do this, that
only he would look at it.”5
Worst of all, at least from a sales point of view, was that there wasn’t a
hint about his personal life—nothing at all about his day-to-day existence, his
struggles with ALS, his wife or his children, his youth. Hawking wanted the
book to be all about physics, not about him. “The material was kind of
underwhelming. You had to see the potential in it,” Guzzardi says. “I mean,
everybody could see Hawking’s potential.” But Hawking the person wasn’t
what Hawking wanted to write about, and that wasn’t going to change. Even
so, Guzzardi was excited about the possibilities.6
The first step was to figure out the audience. “The question is, you know,
where’s the reader? How do we smooth this material out and aim at a reader
at a certain level, and where is that level? Who is that reader?” Guzzardi
asks. “That kind of decision had to be made, and then moving the material in
that direction,… to leaven it with examples and illustrations and whatever
else we might be able to use to lighten and open the material.” Space wasn’t
a problem; the one-hundred-page manuscript needed to be expanded
somewhat to make it to book size. However, even after Gribbin restructured
the book draft, splitting up some chapters, combining others and reordering
the material to extract a coherent narrative, it still needed a lot of work. “It’s
iterative, iterative,” he explains. “So you got a draft and you… go through it,
and you make all kinds of comments and notes. You say, ‘What about this?’
And ‘I lost you here,’ and ‘I really love this part over here.’ And then the
author will address that. Another draft and you go through that process again.
And he typically goes through maybe three drafts before it gets to the place
where you think, ‘Okay, this is close enough where I can really bear down on
it in a sentence by sentence.’”
That’s the most labor-intensive part: going through it line by line, making
sure every idea is understandable and every sentence is clear. “You push and
push and push for explanations for clarity. Until, you know, until I get it
basically, or mostly get it,” Guzzardi says. “Probably, there are places where
I finally would have had to pull up short when it came to a particle spin,
gluons, and quarks. But I just figured this gives you enough of a sense of how
that plays, so that if you really want to know more, you can dig into some
other books, but we’re close enough.”
Despite Hawking’s famous stubbornness, the physicist didn’t resist the
changes his editor was suggesting. “You have to understand how eager he
was to make this work,” says Guzzardi. “He was highly, highly motivated.…
And if this was going to be part of that process, well, by God, he was gonna
buckle down and do it.” Hawking—and his students—were all in.
Of Hawking’s students, Brian Whitt, not Ray Laflamme, was the one who
helped Hawking the most with the writing of the book. “Yeah, it was Brian,”
Laflamme said. “Stephen probably thought that my French Canadian writing,
that people would be upset that, say, pages 92 through 96 were written by a
French Canadian guy.”
“Stephen looked at his students for help because he couldn’t do it on his
own,” Daksh Lohiya, another student of Hawking’s, who had graduated a few
years earlier, told a reporter. “Brian Whitt, a little bit of Bruce Allen (now,
the director of the Max Planck Institute for Gravitational Physics in
Hanover), a little bit of me—everybody made some contribution.”7
Hawking, Whitt, and the other students wrestled with the structure and the
prose, responding to round after round of Guzzardi’s edits. The book slowly
began to take form, and it was improving markedly at each step. The
beginning of Hawking’s one-hundred-page draft had sounded like the opening
to a high-school essay:
From the start of civilization, Man has asked questions such as “When
did the universe begin?” “What happened before the beginning?”
“Will it have an End?” “Is space finite or infinite?” “What is the
nature of time?” “What is the difference between the future and the
past?” The aim of this book is to explain some of the answers to these
long-standing questions that are suggested by modern developments in
physics and cosmology.8
By late 1986, Hawking’s opening to the book was much more engaging:
There is a story about a well-known scientist who was giving a public
lecture on astronomy. He described how the Earth orbits around the
Sun and how the Sun, in turn, orbits around the center of a vast
collection of stars called our galaxy. At the end of the lecture, a little
old lady at the back of the room got up and said: “What you have told
us is all wrong. The World is really a flat plate supported on the back
of a giant tortoise.” The scientist thought he could deal with the old
lady quite easily. “What is the tortoise standing on?” he asked. “Oh,
no, you don’t catch me out like that,” said the old lady. “The tortoise is
standing on the back of another tortoise, and before you ask me what
that tortoise is standing on, let me tell you that it is standing on the
back of another tortoise, and so on.”
Most people nowadays would find the picture of an infinite tower
of tortoises rather ridiculous, but why do we think differently? Do we
know whether the universe had a beginning, and if so, what happened
before then? What is the nature of time? Will it ever come to an end?
Recent developments in physics suggest answers to some of these
longstanding questions.… 9
The final version of the opening of A Brief History of Time, circa 1988, was
striking even by the standards of professional writers:
The final version, of course, ended with one of Hawking’s most memorable
quotations:
All that hard work had transformed the book. Guzzardi had brought out the
best in Hawking. Even the title was significantly improved; originally,
Hawking had suggested From the Big Bang to Black Holes: A Short History
of Time. Guzzardi countered with the much snappier A Brief History of Time:
From the Big Bang to Black Holes. Initially, Hawking resisted the change.
“What I came up with by way of a defense was in a kind of intuitive moment,
I knew how much Stephen loved humor. And I just said to him, ‘You know,
the difference in my mind is that A Brief History of Time makes me smile,
and A Short History of Time does not.’iii That argument carried the day.” At
one point, Guzzardi even had to keep Hawking from cutting that famous last
line. (“Had I done so, the sales might have been halved,” Hawking later
wrote.13)
By early 1988, the manuscript (title and “mind of God” line intact) was
finalized, all the illustrations were done, the cover was selected, and it was
sent out to the printers. According to Lohiya, in that period of waiting for the
books to be produced, Hawking told Whitt and the other students that the
book might actually turn a profit: “‘I think I should share some of the profit
with you,’ he said. ‘So would you like a percentage of the proceeds or a
downpayment?’” After some initial reluctance, Whitt agreed to the sum of
£500: “500 pounds was a hell of a lot of money because our scholarship was
125 pounds a month,” Lohiya recalled. “Judy Fella, the secretary, gave us the
cheques. We got them cashed, went to the pub, and thought we had gypped the
poor man.”14 iv
Nobody had any inkling what would happen next. “You try to do
everything right. You work your ass off, you work on the manuscript, you
work on the cover, you work on the marketing plan, and then you put it out
there,” says Guzzardi. “Then lightning has to strike. You can have all the
elements, but without the lightning strike, it’s just going to disappoint you.
And yet lightning doesn’t strike all that often.”15
W hen the book came out in April 1988, Peter Guzzardi had just left Bantam
Books for another publisher, Harmony. So someone else had to organize the
recall and destruction of all forty thousand copies of A Brief History of Time,
not to mention the pacification of a pissed-off author.
A few weeks to months before a book goes to market, publishers send out
“galleys” of the manuscript to publications and influential people in hopes of
getting the book reviewed in prominent outlets. The esteemed peer-reviewed
journal Nature, which had published one of Hawking’s most important
papers (and a number of his father’s), received a copy, and the editor there
assigned the review to Don Page, one of Hawking’s former students, who had
since become a physics professor at Penn State. Page, an evangelical
Christian, took issue with Hawking’s theology and had some quibbles about
controversial elements of the book, but told Nature readers that Brief
History “should be acquired by anyone seeking to learn about some of the
latest ideas and speculation in cosmology, and to gain familiarity with the
insights that have sprung from the extraordinary mind of Stephen Hawking.”
Behind the scenes, though, Page had found the book riddled with minor
errors—errors that were obvious to those with rudimentary physics
knowledge. For example, an illustration depicting tracks of subatomic
particles was labeled as a field of stars and vice versa. Page called Bantam
to tell them about the problem.16 v
As Hawking put it, “the changeover at Bantam led to such confusion that
the first printing contained a large number of errors, like photographs and
diagrams being in the wrong places or wrongly labeled. They therefore had
to recall it before it got to the shops and do another printing.” However,
Hawking’s memory of the incident was apparently faulty; the US edition had
already reached shops, as it had shipped for a sale date of April 1. (The
recall notice—which told readers that the corrected version would have a
blue cover instead of the original silver one—hit the newspapers on April 3.
The UK edition was published several weeks later, perhaps accounting for
Hawking’s error.) When Bantam sales reps started calling bookstores to try
to get the books shipped back, as Hawking biographers John Gribbin and
Michael White put it, “to their amazement, there were no unsold copies left.
… According to executives at Bantam, this was the first sign that they were
on to something really big.” Lightning had struck.17
Stephen wasn’t the only one who misremembered details about the recall.
Jane wrote, “The first edition had to be pulped at the last minute because of
the fear of legal action on account of certain aspersions cast in the text on the
integrity of a couple of American scientists.” (Jane found a silver lining in
the recall, because “this misfortune allowed a minor omission to be rectified.
Stephen had dedicated A Brief History of Time to me, a gesture which came
as a much appreciated public acknowledgement, but the dedication had been
left out of the American edition.”) Jane was correct about the dedication, but
the aspersions cast at the integrity of two scientists were not the cause of the
recall. In fact, the offending paragraphs sat unchanged for several months,
much to the chagrin of his insulted colleagues.18
Hawking’s attack had come as a complete surprise to its targets. “I think it
was Paul [Steinhardt] who got in touch with me and said, ‘You should know
about this thing in A Brief History of Time,’” says Andy Albrecht, presently a
cosmologist at the University of California at Davis. Nestled in the middle of
the book was a paragraph that contained serious, possibly career-destroying,
allegations against both Albrecht and Steinhardt. In fact, it had already
caused significant damage to Steinhardt. “[Steinhardt] really tried to protect
me,” Albrecht recalls. “I remember him saying, ‘I don’t want you to be
worried about this. You’re very young, you should be pursuing your own
career and not worrying about this.’ And he said, ‘I’m going to take care of
this.’”19
The allegations stemmed from some work that Albrecht, then a PhD
student at the University of Pennsylvania, had done with Steinhardt, his thesis
adviser, in the early 1980s. It was a time of rapid advances in cosmology,
and Albrecht and Steinhardt had devised a new—important—theory called
slow-roll inflation (more in Chapter 11). But they hadn’t been the only ones
to develop the theory. Hawking knew that Andrei Linde, a physicist then at
the Lebedev Physical Institute in Moscow, had come up with a similar idea a
few months earlier—indeed, Hawking and Linde had bonded while
discussing the idea at a Moscow conference in 1981.
It’s common in physics for two sets of physicists to come up with the
same idea in parallel. Perhaps the most famous example involved the most
celebrated occupant of Hawking’s academic chair, Isaac Newton, who
invented the mathematical machinery of calculus at the same time that
continental mathematician Gottfried Leibniz did. It’s also common for these
incidents to lead to accusations of plagiarism and to cause rifts in the
scientific community; this happened with the Newton/Leibniz fight, which
divided British scientists from those on the European mainland for more than
a generation. When it happened with slow-roll inflation, Hawking was the
catalyst.
After Steinhardt and Albrecht published their slow-roll paper in 1982,
Hawking began murmuring to his colleagues that the two had stolen Linde’s
idea—that Steinhardt and Albrecht had heard Hawking mention it in a
Philadelphia seminar in late 1981, and had quickly written a paper to claim
credit. When Steinhardt found out, he was livid; this was the sort of
allegation that could wreck his reputation. Even though Steinhardt, a newly
minted assistant professor, was greatly outranked by Hawking, who occupied
one of the most prestigious offices in all of academia, he had to fight back.
Steinhardt wrote a letter to Hawking stating that he didn’t recall Hawking
saying anything about slow-roll inflation at the Philadelphia meeting—and
besides, the work had already been underway before Hawking came to town.
Steinhardt even sent copies of some of his correspondence to prove it.
Hawking’s conciliatory response seemed to put the issue to bed; he
apparently accepted that Steinhardt and Albrecht had come up with the idea
on their own. And that was that. Until A Brief History of Time.20
Sometime early on in the writing process—as Hawking added padding to
his first draft—Hawking had inserted the following passage:
The implication was clear, and it wasn’t pretty. Hawking had dredged up a
nasty accusation and given it a second life in a very, very public manner. And
it was already causing damage; Steinhardt reportedly found out about the
paragraph when it disrupted his attempt to get a grant from the National
Science Foundation. Steinhardt had to figure out what to do, and fast—not
just to protect his own reputation, but Albrecht’s.22
Luckily for Steinhardt, his protégé had some vivid memories of that
fateful seminar at Philadelphia’s Drexel University in 1981. “That was the
first time I had met Hawking,” Albrecht says. “And one of the things that was
happening at that talk which Hawking gave was that some of the Drexel
students were videotaping it. And I remember having this fierce reaction,
like, they should just come and listen to the science, but they’re fiddling
around with technology instead.” Albrecht’s annoyance with the Drexel
students in 1981 turned to a sense of gratitude in 1988 when Hawking’s
attack was published. “When Paul got nuts and said, ‘There’s this thing in A
Brief History of Time,’ I said, ‘Well, you know, it was videotaped.’”
Steinhart managed to dig up a copy the videotape, and there was no mention
whatsoever of Linde’s version of slow-roll inflation. Steinhardt sent a copy
to Hawking and to Bantam Books—and told the story to Newsweek, which
was doing a cover story on Hawking and his new book. When the Newsweek
reporter asked for a response from Bantam, the publisher announced that the
offending passage would be “expunged from future editions.”23
Hawking himself didn’t apologize until, trying to appease a cosmologist
friend who was close to Steinhardt, he offered to publicly make amends by
writing a letter to the editor of a physics magazine.24 The letter, which
appeared in the February 1989 issue of Physics Today, read, in part:
That is, any perceived insult was merely a result of misinterpreting what
Hawking had written. “It was very maneuvering. And then he mentions a gap
in the tape from Drexel,” says Albrecht. “It’s almost in the league of how
Rosemary Woods might have accidentally erased the [Watergate] tape.… It’s
just this tiny, tiny gap in the flow.”vi He shakes his head. “Face. Trying to
save face.”26
Over the years, Hawking occasionally, slyly, alluded to the incident in a
way that made clear precisely where he stood. “[I] will just say that I first
encountered the idea when I visited Andrei Linde in Moscow in October
1981,” he told a group of cosmologists at a birthday celebration-cum-
physics-conference in 2008. “Again I’m not going to stir up a hornet’s nest by
trying to assign credit for this. I leave that to the Nobel committee.”
Steinhardt’s reputation never recovered in certain circles, such as Hawking’s
beloved Caltech (where Linde landed after the fall of the Berlin Wall). “I
don’t have a very kind view of Steinhardt as a scientist,” says Daniel Z.
Freedman, a supergravity theorist at Caltech close to Hawking, Linde, and
Linde’s wife, Renata Kallosh, also a top-flight physicist at Caltech.
However, Freedman quickly adds, “But he’s done some very, very good
things in other fields.”27
“I think the fact that [Hawking] had been pretty mean made me a little bit
wary around him. Who knows what he’s going to do next. So it did have that
kind of impact, but I sort of sought to transcend that as best I could,” Albrecht
says. “The attitude I chose to take was that human beings do this kind of
thing.”
Albrecht was probably unaware that his words echoed his mentor’s
statement to Newsweek more than thirty years earlier, in 1988. “Hawking is
an outstanding physicist,” Steinhardt had said. “But he’s not a god. He’s a
human being.”28
Day after day, A Brief History of Time sold copy after copy. Not since
Einstein had a theoretical physicist been the sort of person who would merit
a profile in Time and the cover of Newsweek (back in the days when those
magazines were extremely influential). Nobody in the publishing world had
seen lightning like this before. It was getting easier to mistake Hawking for a
god, or at least a rock star.
In Chicago, two superfans, Susan Anderson and Bill Allen, printed five
hundred T-shirts bearing the words “Stephen Hawking Fan Club”—and they
instantly sold out. So the pair printed more. Within two months, the number of
shirts on the street had skyrocketed to eight thousand, and they were
appearing all over the city. Anderson and Allen started getting requests for
shirts from all around the world—including from a certain physicist in
Cambridge (who wanted them in medium and large sizes).29
At the peak of the craze, one Chicago high-school senior admitted that his
T-shirt confused his classmates. “My friends look at the shirt and ask, ‘What
rock group is this Hawking in?’” he told People magazine. “Worse, I have
friends who claim they have his latest album.”30
At first, Hawking couldn’t quite fathom how famous his book had made
him—or what his fame would mean. According to Laflamme, the winter after
A Brief History of Time first came out, Stephen decided to give a series of
eight lectures to undergraduates at Cambridge based on the book, one lecture
for each chapter.
On the day of the first lecture, Laflamme picked Hawking up at the house,
but Hawking was in a rotten mood. “Stephen was very unhappy, and he was
grumpy. Everything I was doing was wrong, so I stopped and I said to
Stephen, ‘I don’t put you the right way in the chair, or it’s too late to go to the
bathroom, or the tea is too hot or too cold. What is wrong?’” Laflamme
recalls. “Peevishly, he looked at me and says, ‘I’m worried about my
lectures.’ He says, ‘I’m worried that nobody will show up.’”31
Soon, it was time to go. The two physicists made their way across
campus to the lecture hall. “You roll from the back door into the guts of the
building, which doesn’t have too many stairs, and we arrived in the room and
it was packed. Packed with people. People sitting on the stairs, probably
breaking all the rules for safety,” says Laflamme. “And suddenly Stephen has
this big grin—that smile. That tells you that even he didn’t expect to catch
that fire.”
For better or for worse, Hawking was a celebrity, recognized not just by
the Cambridge locals who regularly came within a hair’s breadth of running
him over as he wheeled himself heedlessly into the street. He was suddenly
in demand—at galas, for lectures, at festivals. By day, he was feted with
champagne and salmon, which his nurses would carefully feed him, and he
would spend the night spinning maniacally around the dance floor in his
chair. (“He was reckless, driving his wheelchair around and dancing at
parties with it,” says former student Raphael Bousso.) He famously used his
chair as a weapon, allegedly crashing into people who walk too slowly in
front of him, ramming cars that block his ramp, and running over the toes of
people he didn’t particularly like, including Prince Charles. (“That’s a
malicious rumor,” Hawking told a reporter in 2000. “I’ll run over anyone
who repeats it.”)32
Now, after the publication of A Brief History of Time, he had access to
the toes of the richest and most famous people in the world. Laflamme
remembers one infamous visit by a film star: “Someone comes into my
office, and said, ‘Shirley MacLaine is coming, and Stephen is worried,’” he
says. “‘He doesn’t want to be alone with her. Can you come to lunch with
us?’ So, as a student, free lunch, OK? I had no idea who she was, and that
was probably a good thing.” Laflamme rounded up a small handful of
students and went to a little restaurant near the department to join Hawking
and await MacLaine’s arrival, and waited, and waited. “She finally arrives,
and she goes on about how lucky she is to meet this great man, and touching
his hand, and I can see Stephen is trying to stay away from all of this,”
Laflamme continues. “And then she’s there talking about the mind, energy,
and all of this, and all the students, we’re rolling our eyes.” One of the
students got the bright idea to take a spoon, and quietly bend it under the table
and pass it to his fellow students. “So the students suddenly bend spoons. I
remember she was totally, totally over the top.” Jane’s biggest complaint was
that MacLaine was rude: “She spoke only to Stephen and largely ignored the
rest of us.”33
Just as fame brings fans, it brings detractors, too. Even as some reviewers
praised the clarity of Hawking’s prose, others declared it unreadable. Even
the UK publisher of the book, Mark Barty-King, admitted at the time that the
book was “one which I personally found quite difficult to read because of the
subject matter, but one which I considered to have enormous appeal.” In
literary circles, Hawking’s magnum opus quickly became known as the “most
unread bestseller of all time.”34
Hawking was well aware of the claims that purchasers of Brief History
“don’t read it: They just have it in the bookcase or on the coffee table,
thereby getting the credit for having it without taking the effort of having to
understand it. I am sure this happens, but I don’t know that it is any more so
than for most other serious books, including the Bible and Shakespeare,” he
wrote. “On the other hand, I know that some people at least must have read it
because each day I get a pile of letters about my book, many asking questions
or making detailed comments that indicate that they have read the book, even
if they don’t understand all of it. I also get stopped by strangers in the street
who tell me how much they enjoyed it.” That line of criticism didn’t really
sting at all. But another one cut deep.35
In October 1998, New York Magazine ran a long story about the
publishing industry that turned into a vicious attack on Bantam:
PLAYBOY:… Can you tell us a little about your early life, before the
secrets of the universe caught your interest?
HAWKING: Yes. I was born on January eighth, 1942, three hundred years
to the day after the death of Galileo. I was born in Oxford—even
though my parents’ home was in London—because Oxford was a good
place to be during the war.
PLAYBOY: Galileo was tried and imprisoned for heresy by the Catholic
Church for his theories of the universe. Did he have something in
common with you?
I was born on January 8, 1942, exactly three hundred years after the
death of Galileo. However, I estimate that about two hundred thousand
other babies were also born that day. I don’t know whether any of
them were later interested in astronomy. I was born in Oxford, even
though my parents were living in London. This was because Oxford
was a good place to be born during World War II.41
He told this story numerous times over the years in speeches and interviews
as well as in movies about him. It appeared almost word for word in his
2013 autobiography, and in the book he published posthumously in late 2018.
This little anecdote was the public relations equivalent of a shark—it was so
uncannily efficient at its task that it no longer needed to evolve.42
Hawking had been adamant about A Brief History of Time being about
physics rather than the physicist, but he realized that the media—and by
extension, the public—wanted to learn more about the person. So when a
producer picked up the movie option for his book, it caused a bit of a
dilemma, not just for Hawking, but for the moviemakers.
“It’s like any producer, you have a property and you want to turn it into
something. I don’t think anybody had any idea of what to do with it,” says
director Errol Morris. “And I don’t think I had any idea of what to do with it
at first. It was only later, not so long after, but later that I decided that I could
see the movie, that I can see what can I do to make it.”43
Steven Spielberg signed on as the executive producer—the person
holding the purse strings—and soon settled on a director. Errol Morris had
directed a handful of critically acclaimed documentaries, including The Thin
Blue Line, which led to the release of a man who had been wrongly
convicted of murder and put on death row. But what made Morris an inspired
choice was that he had trained to be a historian and philosopher of science.
He was not just fascinated with how humans acquired knowledge and sought
truth, but also, like Hawking, was repelled by certain elements of modern
philosophy of science. (At Princeton, Morris had so frustrated Thomas Kuhn,
the eminent philosopher of science, that Kuhn had chucked an ashtray at him.)
Yet Morris well knew that the real story of Hawking couldn’t be solely about
physics. As he recounted in 2019:
This was not the approach that Hawking initially wanted. “I mean, Hawking
was annoyed with me at times, because he kept saying that it should be really
about the science,” Morris says. “But I said it’s an adaptation of the book.
And in the end, he liked the movie.”45
Jane was harder to convince. “I tried very, very hard to get Jane Hawking
to be in the movie and failed,” Morris recalls. He went to dinner at the
Hawkings’ house, and happened to mention that he played the cello.vii And
they promptly brought out a cello and some sheet music—the Fauré Elegy—
and handed it to the surprised director. “The Fauré Elegy isn’t impossibly
difficult. It’s not easy, but it’s not impossibly difficult. And if I had been
practicing, that would have been okay, but I hadn’t been practicing,” he says.
“I said this to Stephen one time, that if only I had played a little better, she
would have agreed to be in the movie.”46 viii
The movie could survive without Jane, especially since many of
Hawking’s friends and other family members decided to participate,
including his mother, Isobel. But the show could not go on without Steven
Spielberg, who was also extremely unhappy with Morris’ approach.
According to Morris, Spielberg apparently envisioned the film as being more
like Ray and Charles Eames’ Powers of Ten, which was intended to instill in
a viewer an awe for the natural on various scales, from the subatomic to the
supergalactic. But for the intervention of Spielberg’s partner, Kathleen
Kennedy, Spielberg would have fired Morris. In the end, Spielberg merely
wound up pulling his name from the movie credits.
It was an expensive movie for the time—$3 million—and especially
expensive for a documentary. Morris did many of his interviews at the sound
stage in Elstree Studios in England so that he could capture audio of
sufficient quality. But that meant building a replica of Hawking’s office on
the set. Morris’ production designer, Ted Bafaloukos, constructed a perfect
facsimile, down to the smallest details, including the Marilyn Monroe poster
on Hawking’s wall.
During the filming, Morris wound up building a “dictionary” of different
shots of Hawking that could be edited with the voiceover added, thanks to a
duplicate of Hawking’s speech synthesizer, which Hawking allowed the
director to use. Morris’ shots—a close-up of Hawking’s eye, of his hand
squeezing the clicker switch that controlled his computer, of his visage
reflected in the screen of his computer monitor as it flickers with words of
incipient speech—extracted the greatest possible dynamism out of a mostly
immobile human being. And in the end, Hawking was thrilled with the
outcome (he thanked the director for making his mother a movie star), and so
was Morris. “An interview is good if somehow it captured the complexity of
a person that you were interviewing, that on some level—not totally—but on
some level, you created a complex portrait,” he says. “And I do believe A
Brief History is a complex portrait and does capture what I think is truly
interesting about the book.” Even thirty years later, Morris looks back fondly
at the movie. “I don’t always like my work. Often I don’t like it. But I do like
[Brief History],” Morris says. “And I adored Stephen Hawking. I don’t like
most people, but he was something else.”47
Supernova. This was the moment when the star ignited into a blaze of light
that outshone the rest of the galaxy. Hawking’s runaway best seller
transformed the scientist into a public figure—one with a backstory that
turned him into a symbol beloved by billions across the globe. Before, there
had been only an extremely clever scientist working on profound and arcane
questions. After, there was a scientific celebrity the likes of which hadn’t
been seen since Einstein. Before this moment, one needed to understand
Hawking through his physics; after, his physics was almost totally obscured
by his celebrity.
And what a celebrity it was. By the fall of 1988, Hawking was a
superstar, not just in the English-speaking world, but internationally. Crowds
surrounded him wherever he went. When he flew to Spain in October for the
launch of the Spanish-language version of Brief History, he was engulfed in
a sea of people who spontaneously broke into applause as he wheeled by.
“He was good at staying at the center of attention,” says Raphael Bousso, his
student a number of years later. “Everything he said was recorded, and was a
headline in the paper. And you know, he enjoyed that. He enjoyed that very
much.”48
Jane, who accompanied Stephen to Spain along with their youngest son,
Tim, was having a harder time coping with celebrity. “So much sudden
attention was gratifying and disturbing at one and the same time,” she later
wrote. “I felt ill-at-ease in the public eye, conscious of the way I walked or I
held my head or even at the way I smiled. However, the crowds and the
cameramen were not very interested in Tim or me, often pushing us out of the
way in their eagerness to film Stephen, so it was not difficult for us to blend
unnoticed into the throng.” Even at home, Jane felt she had nowhere to escape
to. Since the tracheostomy, there had been team of six or more nurses
attending to Stephen around the clock, and that alone was a big disruption to
the peace of the household. Now that the media wanted access—and Stephen
wanted to grant it—this, too, had become a source of tension. “It was bad
enough having nurses in the house all the time: with television cameras and
reporters as well there would be no privacy for anyone anywhere,” Jane
wrote. “My arguments cut no ice. Quite the contrary, they were represented
as yet further evidence of my disloyalty to the man of genius.” By the middle
of November, Jane had picked out a property in France where she and
Stephen and the children and Jonathan could, “once again, find the unity and
harmony, the modus vivendi” out of the harsh glare of the media spotlight.49
It was not a normal modus vivendi. In the late 1970s, Jane struck up a
friendship with a choirmaster, Jonathan Hellyer Jones, whose wife had
recently died of leukemia. The relationship rapidly grew—from giving the
young Lucy piano lessons to staying after dinner and helping around the house
to something much more.
Stephen had consented. “I would have objected,” he later wrote, “but I
too was expecting an early death and felt I needed someone to support the
children after I was gone.” Jane put it in slightly different terms: “He almost
seemed glad that there was someone else who could relieve him of the
burden of my emotional insecurities so that he could get on with the more
important business of physics.” Regardless, Jonathan had become a key
support for Jane, indeed, a regular part of the family, and someone who
would almost certainly formalize his role when Stephen finally succumbed to
his disease. It was a delicate and sensitive arrangement that required a great
deal of discretion not to embarrass Stephen. Even so, a number of people
who knew the Hawkings at that time relate how awkward it was to have
dinner not just with Stephen and Jane but also with Jane’s lover at the table.50
ix
When Stephen had his tracheostomy in 1985, this modus vivendi was
seven years old, having lasted longer, almost certainly, than anyone expected.
But the tube newly inserted into in his throat meant that Stephen needed
twenty-four-hour nursing, and the cadre of nurses—and, to Jane, their “prying
eyes, listening ears, and gossiping tongues”—upset the fragile balance of the
household. Jane perceived that she and her children Lucy and Tim had
become “second-class citizens as if we, the family, were the lowest of the
low, crouching on the bottom rung of a ladder, at the top of which the angels
—the Florence Nightingales—administered to the deity—the master of the
universe,” she wrote. “In between there were the several echelons of
students, scientists, computer engineers, all of whom were obviously more
important than we were.” Jane soon had recurrent nightmares of being buried
alive.51
The nurse wielding the biggest shovel was Elaine Mason. Mason had met
the Hawkings through her husband, David Mason, who had helped engineer
the computer system and speech synthesizer on Stephen’s wheelchair. A
professional nurse, she joined the nursing rotation, and Stephen soon
preferred her company over the other nurses. The attachment was mutual.
“Elaine volunteered to travel with Stephen at every opportunity,” Jane wrote,
marveling that Mason didn’t feel a “tremendous wrench” when she left her
husband and children at home for such long periods. Even though—perhaps
because—Elaine’s personality was so markedly different from Jane’s,
flamboyant and volatile rather than steady and reserved, Stephen kindled a
romance with the nurse. Jane later wrote that she wouldn’t have begrudged
the two a physical relationship, especially given her own situation with
Jonathan, so long as the pair were discreet—and so long as Elaine posed no
threat to her relationship with Stephen. But, over time, Jane had come to
believe that Elaine was undermining the marriage. “In public and at home,
she seemed to be busily usurping my place at every opportunity, sometimes
apeing me, sometimes undermining me, often flaunting her influence over
Stephen,” she wrote. “She had an unassailable stranglehold over the nursing
rota and had so successfully ingratiated herself that all remonstrance was
useless: any comments would be reported back to Stephen and I would be
castigated for my interference.” And, she continued, her complaints to the
Royal College of Nursing were of no avail.52
Even before the thunderbolt of fame struck, Jane and Stephen’s
relationship was under great strain—strain that even began to affect the staff
at Cambridge University. After the book came out in April 1988, though, vast
fissures began to appear in public.
These cracks are apparent in the 1989 BBC documentary Master of the
Universe. On the surface, the film is upbeat, and Stephen hits all the notes
that so endeared him to the public: his genius, his perseverance, and his
humility. The film ends with a voiceover from his speech synthesizer as he
and Jane tuck Tim into bed. “I have received a lot of help. This has meant
that my disability has not affected my outlook,” he said. “I have a beautiful
family, I have been successful in my work, and I have written a best seller.
One really can’t ask for more.”
Off script, the camera captured a marriage nearing its end. Some of the
details were subtle; before departing the house in the morning, Jane doesn’t
lean down to give Stephen a kiss, but instead tousles his hair and gives him
two pats on the head, almost as if he were an Irish setter rather than a spouse.
Some were blindingly obvious. Jane says that she felt excluded by the
awards and honors Stephen was receiving—that they created distance
between Stephen and his family. “I’m not an appendage of Stephen’s as I very
much feel I am when we go to some of these official gatherings,” she
explained, seemingly on the brink of tears. “I mean, sometimes I’m not even
introduced to people, I come along behind.” Perhaps the most striking
moment in the documentary was when Lucy—then a lively young girl nearing
the end of high school—addressed the cameras: “I’m not as stubborn as him.
I don’t think I’d want to be that stubborn,” she said. “And I don’t think I’ve
got quite his strength of mind, which means he will do what he wants to do at
any cost to anybody else, which I suppose you have to have, really, in his
position.”
The very moments of Stephen’s greatest honors had become dangerous
stressors for his family, and midsummer of 1989—a bit more than a year
after A Brief History of Time hit the market—delivered a one-two-three
combination of honors in quick succession. The last would set into motion a
series of events that led to the end of his marriage.
On the 15th of June, Stephen was to receive an honorary degree from
Cambridge University. It was extremely rare for Cambridge to honor one of
its own in such a way, and it was a major event. Prince Philip, the Queen
Consort, was in attendance in his role as chancellor of the university. (As
Jane writes in her memoir, the prince asked Hawking about his wheelchair as
he rolled by. “Self-propelled, is it?” the prince asked. “Yes,” Jane replied.
“Watch out for your toes!”)53
An even bigger honor came the next day. Every year in June, the queen
announced a list of people who had been granted knighthoods and various
other honors that she saw fit to bestow upon her citizens. On June 16, the
world learned that Stephen Hawking was the newest “Companion of
Honour,” one of some threescore luminaries in the British Empire who had
made distinctive contributions to the arts, sciences, or government. Jane and
Stephen would have a royal audience at Buckingham Palace a few weeks
later. When the couple presented a thumb-printed copy of A Brief History of
Time to the queen, Jane later wrote, the queen asked a baffling question:
“Was it a popular account of his work such that a lawyer might give?”54 x
The third honor came on the 17th of June with a benefit concert in
Stephen’s honor. Jonathan was the conductor of a music ensemble, the
Cambridge Baroque Camerata, which had recently lost its sponsorship. In
May, Jane realized that if Jonathan put on a concert, he could simultaneously
pay tribute to Stephen, raise money for charities such as the Motor Neurone
Disease Association, and raise the profile of the Camerata in hopes of
finding a new sponsor. Stephen agreed, but on the day of, something about the
concert conducted by his wife’s intimate companion had stuck in his craw.
Stephen, Jane writes, “was edgy and disgruntled. His perceptions of the
event seemed to be coloured by the grudging view that Jonathan and the
orchestra had obscured his share of the limelight.” In the argument that
followed, Jane asserted, Stephen pointedly reminded her that since the
“Companion of Honour” was not a hereditary title, she “had no part in it.”55
As soon as she could, Jane retreated, along with Tim, to the new home in
France, where she could “hide [her]self away from the tyranny of the outside
world.” A few weeks later, hoping to defuse the situation, she invited not just
Jonathan and Stephen, but the whole Mason family—David and Elaine and
their children—to France. As Jane put it:
But the modus vivendi was broken. Stephen and Elaine gazed with increasing
antagonism at Jane and Jonathan. The only thing everyone could agree on
seemed to be their universal lack of empathy regarding David Mason, who
was finding himself the fifth wheel in his own marriage. (David was still, to
some extent, in thrall of the physicist. “If he raised an eyebrow, you would
run a mile,” David later told People magazine. “He uses people.” Even after
Elaine left him for Stephen, he would continue to service Stephen’s
wheelchair.)57
Stephen made it quite clear that he wasn’t enjoying his stay in the French
countryside, and tensions rose. Within days, an argument over how Stephen
had treated one of the nursing staff escalated and blew the lid off of all the
resentments that had been building up over the years. “Flames of
vituperation, hatred, desire for revenge leapt up at me from all sides,” Jane
later wrote. Stephen and the Masons returned to England, leaving a shell-
shocked Jane hunkering down in France, waiting for her “lost child” to come
to his senses. However, when Jane finally returned to England in September,
Stephen announced that he was going to move in with Elaine.58
For so long, Jane and Stephen had kept their marriage together. It had
survived as the spousal affection between wife and husband had slowly
transformed into something more maternal. It had even survived as Stephen’s
replacement waited in the wings and began to subsume some of his spousal
and fatherly duties.
It had also survived their disagreement over religion. Though Jane and
Stephen had deep differences in their philosophy and theology, that situation
was little changed since the beginning—with, perhaps, the sole exception that
suddenly people were taking his thoughts on God much more seriously. Jane
herself seemed to view Stephen’s atheism as a primary reason for the
breakdown of her marriage, as she was more than happy to make known.
Journalist Bryan Appleyard later wrote that he was shocked by Jane’s
willingness to air her grievances during a 1989 interview, when she “started
fiercely criticising her husband before I could even turn on my recorder. She
was religious and he was not. In fact, he was aggressively anti-religious, and
his anger with her faith was becoming intolerable. He would not be in the
same room as her devout friends.” Even so, it was Stephen, not Jane, who
sought to dissolve the marriage, and there’s little evidence that, to Stephen,
Jane’s religious beliefs warranted anything more than slight amusement
coupled with a mild dose of condescension. Jane, who felt the religious
schism much more deeply, would have been content keeping the union intact.
What’s more, the woman Stephen next married, Elaine, was just as religious
—and even more demonstrative about her faith—than Jane ever was.59 xi
No, it was not religion that killed their marriage, but the fact that the
Hawking household became increasingly Stephen-centered, first as a
seemingly endless parade of nurses devoted themselves solely to the welfare
of their patient, and then as journalists, celebrities, and adoring fans
prostrated themselves before the newly consecrated god of best-seller
lightning. It’s no coincidence that the bond between Stephen and Jane broke
at precisely the moment that Stephen achieved his biggest triumphs.
The term “triumph” comes from a singular honor bestowed upon a
victorious general by the Senate of ancient Rome. There would be an
enormous parade through the streets of the city, with the conquering general
—the imperator—taking pride of place in a special chariot drawn by four
horses. In the chariot with the general, standing slightly behind him, rode a
slave who held a golden crown above the general’s head. But that slave had
another task as well. As the triumphal procession wound its way through the
cheering crowds, he would periodically whisper in the imperator’s ear:
“Look behind thee; thou art mortal.”60
As Jane told Appleyard in 1989, her role was no longer to be a caretaker
for Stephen in his illness, but “simply to tell him that he’s not God.” Yet of
all people on the planet, the one most conscious of his own mortality was
none other than Stephen Hawking.61
Footnotes
IGNITION (1981–1988)
Behind St. Peter’s Basilica, where tourists aren’t allowed, the chatter of
crowds gives way to the squawking of parrots. The grounds of the Vatican
are green, overwhelmingly green, with immaculate gardens and fountains and
trees of all varieties. Perched atop a small hill is a squat little building that
looks almost like a pagan temple—complete with a statue of the goddess
Cybele. Inside, the walls are adorned with quotations from Seneca and
Cicero. It is almost gaudy in its opulence, with bright frescoes covering the
walls and ceiling and almost every unpainted surface bedecked with marble
of every color known to humanity.
This is the building where luminaries from around the world gather every
other year to advise the pope on matters scientific. Like a number of other
countries, the city-state of the Vatican has a scientific advisory board—the
Pontifical Academy of Sciences—and in 1981 Stephen Hawking was a guest
of the board (and would be appointed a member a few years later).i
This time, the meeting was about cosmology, and the subject of
Hawking’s talk was a wee bit touchy: “the possibility that space-time was
finite but had no boundary, which means that it had no beginning, no moment
of creation,” the physicist recalled in A Brief History of Time. “I had no
desire to share the fate of Galileo, with whom I share a strong sense of
identity, partly because of the coincidence of being born exactly 300 years
after his death!”1
Hawking was joking, but at the Vatican meeting, the physicist proposed
something so bold and radical—and controversial—that it might have made
the pope rather uncomfortable. For the so-called no-boundary proposal,
Hawking argued, dispensed with the need for a creator, for a prime mover
who set the laws of the universe in motion. There was no need for a
watchmaker to fashion the clockwork of the cosmos. There was no need for
God.
Hawking thought the no-boundary proposal to be one of his greatest
contributions to physics. Other scientists aren’t quite so certain; even after
decades of work on it, neither Hawking nor his collaborators managed to
convince the wider community that it was worthy of much attention. But the
no-boundary proposal had a certain elegance, a beauty that was itself
convincing to those inclined to believe that the solutions to fundamental
problems must themselves be beautiful—and radical.
In the early 1980s, Hawking was bursting with optimism about the field of
cosmology. There were a bunch of brand-new ideas about the very early
universe—about the first few fractions of a second after the Big Bang—and
Hawking had been instrumental in giving them life. He was most excited
about his own pet theory, the no-boundary proposal, even though it was
viewed skeptically by most of his peers and almost impossible to explain for
a lay audience. Yet Hawking decided to try. For Hawking was deciding to
write a book about physics. He was bubbling with excitement about his
cosmology and work on black holes, and he hoped that his enthusiasm would
translate into a best seller.
Hawking had his eye on a much bigger audience than his Cambridge
University connections could give him: he wanted to be widely read. More,
he wanted to be understood. It was a drive almost as great as his drive to
understand.
However, as Hawking knew all too well, time was not on his side; the
mere effort of putting words down on paper was excruciatingly difficult for
him. And in 1985, as he lingered in a coma, it seemed for a time that he
would never get a chance to acquire the wider audience that he always
wanted.
Of all of Hawking’s work, the no-boundary proposal is perhaps the hardest
for a layperson to grasp, even in thumbnail, which makes it all the more
remarkable that it was so central to A Brief History of Time. It’s wrapped up
with the baffling idea of imaginary time and the counterintuitive mathematical
formalism that physicist Richard Feynman used to understand the behavior of
quantum-mechanical objects. It brushes up against philosophical problems
that are swept under the rug in the ordinary thinking about quantum theory.
Even the idea of the “boundary” in the no-boundary proposal is not quite
what a non-physicist would think.
To a cosmologist trying to gain a deep understanding of the universe, it’s
not enough to know the laws of physics. An abstract collection of rules,
important though those rules might be, doesn’t necessarily tell us much about
the universe we live in, any more than knowing the rules of chess tells you
what a particular game is going to look like, or whether the person playing
black or white is more likely to win. The rules of the game are only one part
of a full mathematical model of the universe.
The other part is what physicists refer to as the boundary conditions, a
description of a physical system at a given point in time. The laws of nature
take the conditions at the boundary and transform them over time; the system
evolves according to the rules of physics into something new. It’s the
combination of laws and boundary conditions that give physicists a
mathematical model of our specific universe rather than an abstract Platonic
set of equations that might or might not mean anything special.
For example, the laws of gravity apply to all sorts of different
configurations of matter in a given solar system. It doesn’t matter whether the
system has one big planet, or seven thousand little ones, or is composed of
nothing but dust; the fundamental laws are the same in each case. But if you
want to ask a question about our solar system—something as simple as
“Where am I now?”—there’s no way for any mathematical model to yield an
answer without boundary conditions: the positions of the Earth and the other
bodies in the solar system at a given time. It could be the positions of those
bodies a few seconds ago, in which case, the question would be really easy
to answer. It could be the positions of those bodies a million years ago; it
would take some effort to calculate the motions of all those bodies over the
course of a million years, but with precise enough information and a careful
enough calculation of the motions dictated by the laws of nature, physicists
could derive the positions of the bodies in the solar system with high
accuracy. The boundary conditions could even dictate the positions of those
bodies a billion years in the future; the calculations work backward just as
they do forward, so physicists could figure out where we are now based
upon the boundary conditions a billion years hence. Once you have a
mathematical model—laws plus boundary conditions—you can roll it
forward or backward in time as much as you want, answering any sorts of
questions you can ask of such a model. But the boundary conditions are as
important to the model as the underlying physical laws are.
This poses a particular problem for cosmologists who are trying to come
up with models of the very early universe. It’s not just that they don’t know
the underlying physical laws that were operating in the very dense, hot
conditions shortly after the Big Bang—that’s the realm of quantum gravity,
which hasn’t been worked out yet. But even with a grand unified theory in
hand (and, at the time, Hawking was fairly confident that physicists would
build one by the turn of the century), that wouldn’t be enough to build a
mathematical model of the cosmos that, as Hawking put it, “would account
for the whole universe at one go”:
It now seems possible that we might find a fully unified field theory
within the not-too-distant future. However, we shall not have a
complete model of the universe until we can say more about the
boundary conditions than that they must be whatever would produce
what we observe.2
Perhaps the best place to start the story of the no-boundary proposal is in
the late 1940s, when Stephen Hawking was only six years old. Richard
Feynman was then a young physicist, having recently left bomb work at Los
Alamos for more academic pursuits. And he was hard at work wrestling with
the fundamental puzzles of quantum mechanics.
According to the rules of quantum theory, subatomic particles behave in
ways that violate the laws of motion that we’re used to—worse, they violate
common sense. It’s impossible for a brick or a cat to be in two or three or
four places at the same time. However, protons and neutrons and electrons do
this all the time. Indeed, the laws of physics describe such particles as being
in a state of “superposition”—doing several mutually contradictory things
simultaneously—as a matter of course. If you throw a Ping-Pong ball at a
screen with two holes in it, the ball must go through the left hole or the right
hole (or neither); it can’t go through both at the same time. Shoot an electron
at a barrier with two holes, and unless you explicitly prevent it from doing
so, it will automatically zoom through both the left and the right at the same
time.
As strange as this behavior might seem, the mathematics of quantum
mechanics makes perfect sense. If you want to find out what happens to an
electron being shot at a two-slit barrier, all you have to do is set up the
equations of quantum theory in the right manner, and out pops the answer,
encoded in a mathematical object known as a wavefunction. That
wavefunction—so named because the quantum-theoretic equations treat
objects almost like they’re waves—contains all anyone could possibly want
to know about the electron. (In fact, it contains more than anyone could
possibly know; thanks to uncertainty, an observer won’t be able to obtain all
the information from the wavefunction, just some of it.)
That’s all very nice and good for a physicist who needs to do
calculations. The answer always comes out correct. But to Feynman, getting
the correct answer wasn’t sufficient. He wanted a deeper understanding of
what was happening to the electron, and he didn’t feel that he really
understood something until he could visualize it. And he couldn’t really
visualize a wavefunction, whatever that wavelike object might be, whether it
was passing through two slits or doing something more complicated. So he
had to come up with another way of describing quantum objects, a way that
could be seen in his mind’s eye. As he told an interviewer,
The electron does anything it likes. It just goes in any direction at any
speed, forward or backward in time, however it likes, and then you
add up the amplitudes and it gives you the wave function.5
And since i is the square root of negative one, the i2 kills the negative sign in
front of the τ2. That is, all of a sudden, the distance formula looks like this:
The troublesome minus sign has disappeared—there’s no hint of the strange
Lorentzian geometry that plagues ordinary spacetime. It looks Euclidean, so
long as we use imaginary time in our equations rather than time.
Mathematically, this is no big deal. In some sense, going from time to
imaginary time is roughly equivalent to turning your head sideways—rotating
your view by 90 degrees—in a complicated higher-dimensional space. (In
fact, scientists call it a Wick rotation after the person who first used the
technique, Italian physicist Gian-Carlo Wick.) This substitution-cum-rotation
does, in fact, make the mathematics much, much easier. But philosophically,
it means that the equations no longer describe the motions of objects in the
universe in the ordinary way. Watching something moving “forward” in
imaginary time doesn’t have the same obvious interpretation as observing
something evolving as ordinary time passes. “So let’s give up on time being
real. Let’s pretend time is imaginary. And now everything’s fine,” says
Turok. “The problem with it is you can never calculate anything in real time
from this Wick-rotated picture, or it’s very difficult to do that.” Physicists are
okay with this loss of interpretability if they gain more in understanding than
they lose. For any non-physicist who wants to follow what’s really going on,
though, it’s a huge barrier to comprehensibility.10
The Wick rotation takes a Lorentzian manifold of spacetime, with all its
strange geometry and bizarre relativistic effects, and transforms it into a
manifold of space-imaginary-time with a super-simple Euclidean geometry
that’s much easier to handle. For all the disadvantages that this device
creates, it has a number of big advantages. For Hawking, the decisive one
was that the Feynman path integral method—which simply didn’t work in
spacetime—was just fine in space-imaginary-time. Hawking could start
calculating the wavefunction of the universe, so long as he could come up
with a consistent way of describing all possible cosmoses to feed into the
Feynman path integral machinery.
In 1981, when Hawking visited the Vatican, his thinking on the subject
was still relatively new, and Hawking was operating mostly on intuition and
hope. But when he was looking for a consistent description of cosmoses, his
intuition was telling him something that was extraordinarily profound, if true.
Hawking’s visualizations of the cosmoses were telling him that all possible
universes, viewed through the lens of imaginary time, were smooth, closed,
finite, and “compact.” They didn’t have singularities, and didn’t extend
forever in any direction.ii And one can prove, mathematically, that a universe
with such properties would have no boundaries—neither in space nor in
(imaginary) time. It would have no beginning and no end.
From a physicist’s point of view, this notion destroys the problem of
boundary conditions at the Big Bang entirely. There can’t be boundary
conditions to the universe because the universe has no boundaries! The Big
Bang isn’t really a boundary, an end, an “edge,” of our cosmos any more than
the South Pole is an “edge” of the Earth. And so in Hawking’s no-boundary
model, the laws of physics are enough to determine everything about the
cosmos. As Hawking put it, “if the universe is in a no-boundary state, we
could, in principle, determine completely how the universe should behave,
up to the limits of the uncertainty principle.”11
Hawking liked to put a theological spin on his idea of a universe with no
boundaries—the idea of a universe with no beginning and no end, to him,
seemed to obviate the need for a creator. In some sense, God was necessary
to set the boundary conditions, to give the universe its beginning or bring it to
an end. But a universe without a beginning or end would just be: it wouldn’t
need to be created or destroyed. Needless to say, Hawking’s interpretation
wasn’t terribly popular with theologians.
The no-boundary proposal wasn’t born as an attempt to banish God; it
was a complicated way of applying a thirty-year-old technique for handling
quantum systems to the universe as a whole. And, as one might expect, given
that it’s an attempt to sum up all possible universes, it rests on a whole bunch
of assumptions and intuitions. Even after he worked with Hartle in the early
1980s to work out some of the kinks in the theory, and labored on the topic on
and off in the decades after that, the no-boundary proposal wasn’t embraced
by most of the scientific community. “The idea has never really been
accepted,” Turok says. “I would say 90 percent of cosmologists or
theoretical physicists who have an opinion… most people don’t even form an
opinion; of those who do, 90 percent of them would say they probably didn’t
agree with it, or they thought there was some problem with it.”12
To Hawking, however, it was a central pillar of his legacy. And it was a
potent metaphor. Quite naturally, the last chapter of his last book—his
autobiography—was entitled “No Boundaries.”13
In the early 1980s, Stephen Hawking’s voice is distinctly baritone, but it’s
almost impossible to understand what he is saying. His larynx and tongue and
lips still have enough control to make sounds that come out with the pacing
and the structure of ordinary speech, but it doesn’t sound like language at all.
He makes a creaky, keening sound with a guttural rumble behind it. It sounds
almost as if a great cat were trying—with infinite patience—to explain
something in English to an uncomprehending audience. So, every few
seconds, he has to break off as one of his inner circle, such as his graduate
assistant or someone else who can interpret his utterances, translates his
words.
Even so, his lectures and chats are frequently punctuated by peals of
laughter. Hawking was somehow able to display his wicked sense of humor
in his talks, just as he could in his writing, even though he had to speak
through a translator, or later, a voicebox, which nearly obliterated any
exercise of comedic timing. “Somehow it worked for him,” explains John
Preskill. “Often, he’d make some sort of comeback, and it wouldn’t be
immediate; you’d know it was coming up… somehow, for him, the delay was
part of what made it funny. Sometimes, he’d immediately say ‘Rubbish!’ That
was one of the ways he would disagree with you. Very British, you know.
You’d say something, and he’d say ‘Rubbish!’” Even though people tended to
treat Hawking with deference, Preskill, by his own account, treated him in a
somewhat snide way. “I’d make fun of him, and he’d like this,” Preskill
recalls. “Whenever we’d get together, I’d be irreverent around him, and
when he would make a comment, I’d say, ‘What makes you so sure about that,
Mr. Big Shot?’ and that kind of thing. And he really liked that.”
And by the early 1980s, Hawking was a big shot, and to a greater and
greater extent, not just in the physics circles he moved in. The press was
beginning to notice the young scientist, just entering his forties, who was so
highly regarded by his peers, and who had such a unique and heartrending
backstory.
Even in the early 1980s, Hawking was no stranger to the media: science
journalists had quoted him in newspapers and magazines quite a few times
over the years as an expert on black holes. But around that time, reporters’
interest in Hawking was beginning to shift from Hawking as a subject-matter
expert to Hawking the human being—he was no longer being used to
comment on a story, but instead becoming the story himself.
In 1981, the New York Times ran a piece by journalist Malcolm Browne
titled “Does Sickness Have Its Virtues?” The article clearly articulated one
major leitmotif in the coverage to come:
When Hawking’s face with its “mild cetacean smile” graced the cover of
the New York Times Magazine in early 1983, readers were treated to a heavy
dose of this physical-infirmity-begets-genius theme:
Lucy Hawking was just turning twelve, and was graduating from Newnham
Croft, the local elementary school. She had gotten in to Perse Girls, a well-
regarded school—but it was one the Hawkings would have to pay for. Thus
was born the idea for A Brief History of Time. “I first had the idea in 1982 of
writing a popular book about the universe,” Stephen later wrote. “The
intention was partly to earn money to pay my daughter’s school fees.”17
While this might have been the main impetus for Hawking to attempt to
write a popular book, it was not his entire motivation. For Hawking wanted
to be well known outside of his community. He was not one of those
scientists who eschewed (or actively scorned) popularizers. In fact, he
wanted to be one himself. Over the years, he had tried his hand at writing a
few articles for popular consumption in magazines such as Scientific
American and New Scientist. The time seemed ripe for a more ambitious
attempt. Carl Sagan’s Cosmos had shown that the market was there, and
Hawking thought he might be able to leverage his increasing popularity in the
press to reach a wider audience than before. Hawking later wrote that the
impulse for writing a book came from the message he wanted to convey:
“The main reason was that I wanted to explain how far we had come in our
understanding of the universe. How we might be near finding a complete
theory that would describe the universe and everything in it.”18
The year 1982 was a high point in Hawking’s excitement about
cosmology. There was a brand-new theory, inflation, that seemed to explain
what had happened in the first few moments after the Big Bang; a year prior,
Hawking was at the center of an attempt to work out the kinks in the idea.
(More about this in the next chapter.) And, of course, after a number of years
of research, he had just unveiled his no-boundary theorem; in 1982, Hawking
was busy hammering out the details with Jim Hartle. And in his boundless
enthusiasm, Hawking—echoing the gospels—was telling his audiences that
some of those present would likely live to see a theory of everything.iii It was
almost inevitable that Hawking would start writing a book to spread this
excitement—and these ideas—to a wider audience.19
Hawking had published a few books before; they were academic works
intended for a technical audience published with Cambridge University
Press. Like most books published by academic presses, they were never
expected to make much money for the publisher, much less for the author.
However, there are always exceptions, and in the mid-1970s, Hawking and
his friend George Ellis had written a very technical treatise explaining some
of the latest developments in relativity theory from base principles, and it
had been very well received by the community. “Now that had done
incredibly well,” says Simon Mitton, who was then the director of science
publishing at Cambridge University Press. “And it meant my colleagues at
the press both in Cambridge and New York were very anxious. ‘Simon, can
you get Stephen to do something for us?’”20
Mitton, himself an astronomer, had been the departmental secretary of
Cambridge’s Institute of Theoretical Astronomy when Hawking worked there
half time around a decade prior; he had gotten to know Hawking quite well,
as his departmental duties had included looking after Hawking’s needs. The
two remained in contact after Mitton accepted his new position at Cambridge
University Press. “I got on with him very well when I had responsibilities for
him at the Institute of Astronomy. And I got on with him well when he
consulted me about publishing matters,” Mitton says. “Stephen’s secretary,
Judy Fella, contacted me. And she said, ‘Stephen would like to talk to you
about a typescript. I’ve just finished typing it, and it’s for a popular book
about the universe.’”
It was the fall of 1983, and Hawking had been working on the typescript
for months—partially cobbled together from previous lectures, partially
novel material—and he was eager for Mitton to take a look. “Before I started
reading it, I said to him, ‘Stephen, what’s your motivation for writing this?’”
Mitton says. “His response was moving, but also clichéd. He said that his
motivation for writing the book was security for his family, and particularly
the education of his three children.” That wasn’t a realistic goal for a book
published with a university press; advances were modest and sales generally
were, too. Even in the trade press, it’s rare for a first-time author to make
enough money to provide security for his or her family. When Mitton
explained that this wouldn’t be likely, at least with Cambridge University
Press, “[Stephen] said, ‘Well, please do me a favor and read it and give me
some feedback, and together we can decide what to do next.’”21
Mitton was eager for a book by Hawking, and after digging into the
manuscript, he thought the general idea of the book was good. Unfortunately,
he found the draft far too technical for a popular audience. Not many years
prior, Mitton had convinced another eminent cosmologist, Paul Davies, to
write a cosmology book for the general audience. Though Mitton had been
very excited about the book’s prospects (and had come up with a striking
title: The Accidental Universe), the press’s marketing director had thrown
cold water on the project. “He said, ‘Simon, this book is unsellable. I hope
you’re not printing too many copies,’” Mitton recalls. “I said, ‘What do you
mean? Paul Davies is world famous! You can ship them by the bushel.’ He
said, ‘No we can’t. There are equations on dozens and dozens of pages and
every one of those halves the market.’” The marketing director was right;
when the book came out in 1982, sales were poor. The last thing Mitton
wanted was a repeat performance.22
So his feedback for Hawking was rather disappointing. Hawking later
told Time magazine, “Someone told me that each equation I included in the
book would halve the sales.” Mitton recalls that Hawking was resistant to
making the book less technical, at least at first: “His initial response was,
‘Well, you say that, Simon, but I’ve only used the kind of mathematics which
seventeen- and eighteen-year-olds have to do in order to satisfy matriculation
and university entrance.’”23
Despite his cavils, though, Mitton was convinced he could turn the book
into a big seller, and he approached the powers that be at the university to get
permission to offer Hawking a contract. “They agreed to the largest advance
that I’d ever paid out to that time.… They agreed on £10,000, half on
signature, half on delivery,” Mitton says. “I knew that wouldn’t be enough to
satisfy him, but he was my friend, and I wanted to demonstrate to him that I
was passionate about the book, and I did the best I could.” Unfortunately for
Mitton, as Hawking no doubt knew, he could get much more from a trade
publisher. For a bit more than a year earlier, in late 1982, Hawking had found
a literary agent. Or, more precisely, a literary agent had found him.
Daniel Z. Freedman was an esteemed physicist and one of the early
pioneers of supergravity—at the time, what Hawking considered to be the
leading candidate for a theory of everything—and he had learned that
Hawking was thinking about writing a popular book. Freedman’s sister was
married to a high-powered literary agent, Al Zuckerman. “Danny asked if I
would like to talk to Hawking, and I said sure, and I went to Cambridge,”
Zuckerman says. By the time Hawking approached Mitton, Zuckerman had
already convinced Hawking that he would be able to get a sizable advance
for the book.24
Zuckerman had very good reason to think so. Back in January 1983—in an
odd coincidence—Zuckerman had had a lunch appointment with an editor at
the Bantam publishing house, Peter Guzzardi. In Guzzardi’s bag was a copy
of The New York Times Magazine, the one with the image of Stephen
Hawking on the cover. “It was just a wonderful article so beautifully written.
And I was just struck by it to the point where I was carrying it around in my
bag,” Guzzardi recalls. “And it was a combination of the ambition [the
author]iv conveyed, the ambition of what Stephen was doing professionally,
this quest he was on, and the contrast with his being trapped in his body, and
that just struck a nerve with me.… Here’s this guy exploring the outer limits
of the cosmos and he can’t tie his shoes.” When Guzzardi mentioned the
article to Zuckerman, “Al Zuckerman said to me, ‘Hey, it’s amazing that you
should mention this because I’m actually on it.’”25
It took months of work, but by the end of 1983, the proposal, which
included Hawking’s one-hundred-page typescript, was ready. Zuckerman
started approaching publishers, including Bantam. Bizarrely—and perhaps
absentmindedly—Zuckerman didn’t send the proposal to Guzzardi, despite
their earlier conversation. “He must have forgotten about that because when
the time came, he sent the proposal to another editor at Bantam, Jeanne
Bernkopf, and she was a brilliant editor,” Guzzardi says. Bernkopf
remembered that Guzzardi was interested in Hawking and passed the
proposal on to him rather than taking it to the editorial board herself. “It was
an act of integrity and generosity and kindness that was rather remarkable in
that kind of competitive environment,” Guzzardi says. So Guzzardi would try
to acquire the book for Bantam. However, Bantam wasn’t the only publisher
out there.
“In New York practically every publisher in town was interested in the
book,” Zuckerman told a trade journal shortly after A Brief History of Time
was published. “I had six-figure offers from about six.” Bantam came in at
$225,000—a very sizable advance for its day, beating the Cambridge
University Press offer by more than an order of magnitude. “Bantam was
hungry for prestige. They had a prestigious success with [Chrysler CEO] Lee
Iacocca’s memoirs, and they liked the taste of that,” Guzzardi says, adding
that he hoped a Hawking success could show that Bantam could “play with
the big boys.” But, as it turned out, Bantam wasn’t the highest bidder for
Hawking’s book.26
W. W. Norton was a publishing house best known for its anthologies,
textbooks, and other academic-type books, but it had a few highbrow
nonfiction successes as well. In its pipeline was a best-seller-to-be called
Surely You’re Joking, Mr. Feynman, a collection of anecdotes about
Feynman’s life and work. If Norton had grabbed the Hawking contract, it
would have become the go-to house for scientists seeking to popularize their
work. But even though Norton had tendered the highest bid, Hawking wasn’t
bound to accept its offer. He could choose whichever publishing house suited
him best.
Guzzardi thinks that one sentence he put in a letter to Hawking clinched
the decision. “He was always hungry, and was drawn to Bantam because of
the promise I made in a letter that we could distribute the book in ways that
other people couldn’t at that point, which was true,” Guzzardi says. “The
sentence that we could put the book in airport bookshops was really the one
that most captured his imagination.” It was a technical advantage that Bantam
had in those days; the publisher distributed its wares through paperback
wholesalers that could put books into drugstores and supermarkets and
airport bookshops. “And they could use that leverage to put their hardcover
books there, too, when appropriate, which is not often but was certainly the
case with Iacocca, for example.” That clinched the deal. Bantam would be
the publisher, at least in the United States and Canada, and Zuckerman
eagerly set out to secure contracts for foreign and translated rights. Even if
the book didn’t sell well, Hawking stood to make a pretty decent profit from
the advance alone.27
Mitton wasn’t insulted when he found out that his friend and colleague
had signed with a big publishing house; after all, good old Cambridge
University Press couldn’t come close to offering the sort of money or
marketing or distribution that Bantam would be able to put behind Hawking’s
work. But Mitton gave Hawking a few words of wisdom:
In addition to the book writing and media attention, Hawking was spending a
disproportionate amount of time being showered with honors (including
being named a Commander of the British Empire by the queen), traveling,
and fighting for disability rights. Stephen and Jane (and Jonathan Hellyer
Jones) helped the newly founded Motor Neurone Disease Association in its
fundraising efforts and fought for access to services for the disabled; when
one of the colleges at Cambridge planned to put in a library without
provisions for wheelchair access, the Hawkings raised a public stink. With
all this activity, it’s a wonder that Hawking was able to get any physics work
done at all.29
However, the distractions from work were welcome, at least to some
extent. For all his enthusiasm about the no-boundary theorem, Hawking was
worried that his great scientific achievements were behind him. “As far as
theoretical physics is concerned, I’m already over the hill. Actually, quite far
over the hill,” Hawking told a reporter in 1982. “Well, you know most of the
best work in theoretical physics is done by people at a very early age—
usually by people in their twenties. So being forty is not a stage in life where
one expects to make great discoveries in theoretical physics.” Luckily for
Hawking, he had his pick of the brightest twenty-something-year-olds that
Cambridge had to offer—young physicists in training whom he could shape,
and who could, in turn, shape his thinking and help him with his work.30
Ray Laflamme won Hawking as a mentor in 1984 thanks to his
performance on the famously brutal Mathematical Tripos test. Not only was
the Tripos long (administered over several days) and fiendishly difficult, but
it was pretty much the only thing that a student was judged by. “I went to the
exam and said, ‘Shit, this is damn hard,’” Laflamme recalls. “Some questions
I just have no clue how to answer them.” And one tradition that Laflamme
found particularly hard to swallow was that the results were announced
publicly, at the Senate House, by a professor—to a gathered crowd, the prof
would read out a list of names of people who had passed and who had failed.
“To me, this is really barbaric,” Laflamme says. “I cannot go there. I cannot
go with my peers and sit and listen to this.” So Laflamme skipped the
readings. When he went to the department in the morning, he was pleasantly
surprised to find out that there was a note for the people who had passed
with distinction to go speak to the head of the group—and his name was
among them.31
Laflamme duly knocked on the department head’s door. “He said,
‘Professor Hawking is waiting for you downstairs.’ I was thinking, ‘Is this
real? Am I in a dream?’ And I would pinch myself, just kind of try to wake
up. Okay, this is really a fantastic dream, but it cannot be true.” But Hawking
was indeed there, along with a nurse, waiting. “‘You’ll be my student and tell
me what you’re up to.’ So he asked me what I was going to do for the
summer,” Laflamme recalls. “He must have thought that I was a completely
dumb idiot because I had a problem putting three words together and being
coherent. Afterward, I learned that people in front of him would kind of
freeze.” Hawking gave Laflamme a tremendous amount of reading to do over
the summer, and in the fall of 1984, Laflamme began working on the no-
boundary proposal and the wavefunction of the universe.
“There wasn’t much of a separation between his personal life and the
students,” says Raphael Bousso, who worked with Hawking a decade later.
“He was very gregarious, and he kept us much closer than I think most
advisers keep their students.” And before 1985, when Hawking’s
tracheostomy forced him to hire nurses to watch him around the clock, some
of his students would take on nursing duties—helping their adviser bathe and
urinate and dress was as much a part of their lives as working on physics
problems. “We would help him go to the bathroom and feed him and do
things which makes a relationship with your [mentor] totally different,”
Laflamme says. “I depend on my students’ brains, not physically. But Stephen
was very strongly dependent on his students.” Yet the students whom
Hawking depended upon never resented their adviser or begrudged him his
needs—quite the opposite. They seemed to treasure the intimacy that their
unusual relationship with their adviser brought.32
Laflamme, for one, particularly enjoyed spending time with Hawking at
airports. Travel for Hawking was quite a production—carting baggage,
assembling and disassembling the wheelchair, carrying Stephen on board and
dealing with his demands—which brought Jane each time to near her
breaking point.
Laflamme also suffered from the drama of travel with the Hawkings, but
from his point of view, not all of it could be laid at Stephen’s feet. “I
remember I’m traveling somewhere. And getting Stephen up in the morning,
dressing him, giving him breakfast, and making ready to go somewhere,
which took up a lot of time,” he says. “And I remember sitting in a car
waiting for Jane to show up. And saying, ‘I applied to be a student in
theoretical physics. I didn’t apply to be part of a Hollywood drama.’ But I
could imagine it was damn hard for her.” But to Laflamme, the rigors of
travel were well worth it, for they provided an opportunity to have
uninterrupted time with a great mind. “Just waiting for a flight turned out to
be these magic moments for me. Because he was stuck there,” Laflamme
says. “So we’re just waiting. And that’s often where I would ask him
questions. ‘Why Stephen, do you think this? Like, why do you think summing
over Euclidean geometry is the right thing to do for the wavefunction of the
universe, like, where does that come from?’”33
And when a student shared Hawking’s irreverent sense of humor—and
fondness for pranks—road trips occasionally were punctuated by surreal
moments of comedy. During one trip, Hawking told Laflamme that he had to
go to the bathroom, so they wheeled over to the men’s room—but it was
closed for cleaning. Laflamme asked the physicist what he wanted to do.
“Then he looked at the women’s bathroom. So I said, ‘You want to go there?
Okay, fine.’ So we go in the women’s bathroom,” Laflamme remembers. And
so the pair go in, and the bathroom is empty. Laflamme pulls out a receptacle
for urine and begins preparing Hawking for the procedure. “And so two
women come straight in and he’s in a wheelchair, his fly open, me with the
bottle, and look at him. And he has this big smile on his face.” Before the
women could register their offense, Laflamme quickly turned toward them.
“And I said, ‘If you want to take my place, feel free!’ And Steve had this big
grin on his face. He was probably happy that the men’s bathroom was busy;
he thought it was a very good joke.”34
Despite his health, Hawking traveled widely and often, with trips timed to
take advantage of an academic schedule. Christmas breaks, spring and fall
breaks, and summers were chock-full of travel, to the point where Jane—
who was attempting to overcome a phobia of flying—complained that “travel
had become an obsession with him and he regularly seemed to spend more
time in the air than he did on the ground.” Summer 1984: Chicago. Fall 1984:
Moscow. Spring 1985: a tour of China, during which Stephen managed to
drive his wheelchair on the Great Wall. Summer 1985: Geneva.35
Travel was central to how Hawking did physics; it was at scientific
conferences that he and his colleagues forged collaborations, identified big
problems to tackle, and even occasionally solved those problems. Travel
tilled the intellectual soil, mixing ideas from the best minds from all over the
world. Different scientific groups had different cultures, different strengths
and weaknesses, and different philosophies. Over and over throughout
Hawking’s career, his thinking was shaped by the people whom he met at
meetings and conferences—Kip Thorne and John Preskill and the rest of the
Caltech crew, Andy Strominger and Sid Coleman in Massachusetts, Andrei
Linde and Alexei Starobinsky and Yakov Zel’dovich in the Soviet Union. For
physics, even theoretical physics, is a deeply social pursuit.
Of course, Hawking’s time away from Cambridge wasn’t entirely devoted
to physics. In the summer of 1985, he took Laflamme to spend some time in
Geneva, Switzerland—a short drive from CERN, the European center for
high-energy physics. The pair were, in fact, working on a physics problem
together, but it had to do with Hawking’s wavefunction-of-the-universe work,
which had almost nothing in common with the work that was going on at
CERN. In truth, Laflamme suggests, Hawking was using Geneva as an excuse
to set up shop on the European continent to satisfy his craving to listen to
Wagnerian opera at the annual Bayreuth Festival, the mecca for Wagner fans.
But the pair did, in fact, have physics to do first.
One of the main features of Hawking’s no-boundary proposal was the
underlying principle that any possible universe was compact, that it couldn’t
extend without limit in time or in space. At the time (until Turok’s work a
decade later, which gave a way out), Hawking thought this meant that the
universe had to end in a Big Crunch: a reverse Big Bang. Just as a Big Bang
is a birth of a universe from a point, a giant explosion that expands rapidly, a
Big Crunch is the death of the cosmos, a rapid contraction of spacetime to a
point and then to nothing.
To Hawking, this presented an interesting dilemma. The Big Bang takes
our universe from a relatively simple state—something that looks roughly the
same everywhere—to something highly disordered and lumpy and
inhomogeneous, with stars and galaxies and clumps of gas scattered through
vast tracts of vacuum. In physics terms, the entropy of the universe increases
as our universe expands. (More on the concept of entropy in Chapter 13.) In
fact, this increase in disorder is the main way that we can tell which way
time is flowing: if we see a film where shards of glass suddenly assemble
themselves into a vase, or cream suddenly separates itself from coffee due to
the stirring of a spoon, we instantly know that the film is being played
backward, because self-assembling vases and self-separating creamers
require entropy to decrease rather than increase.
But if the universe stops expanding and eventually contracts into a Big
Crunch, wouldn’t that mean that the disorder of the universe would stop
increasing? Right before the universe ends, it would have to wind up in a
state as simple as it was right after the universe began—the entropy of the
universe would have to decrease during the contraction, not increase. And if
this is the case, wouldn’t the arrow of time change direction? Would the
universe run backward like a film run in reverse? Would dry bones in their
graves suddenly start growing tendons and skin, and upon being lifted out of
the dirt by backward-walking pallbearers, be infused once again with the
breath of life? At the time, Hawking thought this was the only logical
conclusion, despite the seeming paradoxes it might cause. As he wrote in his
earliest draft of Brief History:
That is, yes, jugs would leap off the floor and people would be coming back
from the dead. But all intelligent life would be wiped out at the moment the
universe begins to contract, ensuring that nobody would be around to
perceive an apparent contradiction.
That was his intuition, anyway. Hawking hadn’t yet done the detailed
calculations to determine whether the arrow of time would, in fact, reverse
like he thought it would. And that’s where Laflamme came in. Sharing a small
office in Geneva with Hawking, Laflamme was trying to work out the
mathematics of what happens to entropy in a collapsing universe for
Hawking. “And he really wanted to know the answer, and it was not going
his way,” Laflamme says. “And every half hour I would hear, ‘Do you have
an answer yet?’ Not exactly the most relaxing. I was so tired. After two days,
I was like a zombie walking around.”37
In the days before his tracheostomy, Hawking didn’t need round-the-clock
nursing to keep him alive. But he still needed attendants to be around him
twenty-four hours a day, because he was all but immobile in his wheelchair.
And that duty fell to Hawking’s nurse, assistant, student, or wife, depending
on who was around at the time. Jane hadn’t accompanied Stephen to Geneva;
she was driving and camping through Belgium and Germany with Jonathan,
her daughter Lucy, and her six-year-old son Timothy, with plans to meet up
with Stephen at Bayreuth later in the summer. For her, it was a welcome
respite from caring for her husband, and a rare opportunity to let the “poor,
sickly plant of [her and Jonathan’s] relationship… come out into the open for
an airing and to blossom,” even though it was “watered with tears of tension
and guilt.”38
It was a heavy burden to care for Stephen, as Laflamme was quick to
admit. After two exhausting days in Geneva, the weekend had finally arrived.
“I can sleep late and just relax,” Laflamme says, smiling. “But Stephen wants
to go and look at the mountains. He wants to go downtown. Here goes the
slave driver.” So Laflamme and Hawking’s nurse and his assistant piled into
the car and drove to downtown Geneva, where Hawking decided he wanted
to visit a record shop. After lifting Hawking up the stairs to get into the shop,
they realized Hawking couldn’t see the records from his wheelchair; all the
displays were for people standing up. So Laflamme and the others had to pull
down large bunches of records and show them to him one by one, waiting for
him to say yes or no before flipping to the next one. It was an exhausting
process, and then, after lifting him out of his wheelchair, getting him down
the stairs, and positioning him in the wheelchair again, Laflamme was ready
to collapse.39
“He’s saying something. The nurse and the secretary have gone
somewhere, so I’m on my own with him. And I cannot understand what he’s
telling me. I hear, ‘Put me off,’” Laflamme recalls. “I said, ‘Stephen, what
are you saying?’ He said, ‘Put me off.’” After a few more repetitions,
Hawking is getting extremely frustrated, and Laflamme is getting flustered,
not knowing what to do. “I look at Stephen, and I said, ‘I’m going to walk to
the corner of the street, and I’m gonna come back. And by that time, I will
have relaxed and hopefully I can understand you.” Laflamme takes a deep
breath and walks to the end of the block and back. “I can see smoke coming
out of his ears,” Laflamme says. “‘Put me off!’ I said, ‘Yeah, you’re telling
me, ‘put me off.’ But then I see his eyes are turning, and he’s looking at his
hand. He wanted me to turn off the wheelchair—‘Turn me off’—so that he
could put his hand in a better way on the wheelchair. Because if he was
trying to move his hand [while the wheelchair was on], the wheelchair would
move.” Everything that Hawking wanted to do, no matter how seemingly
simple, meant that the people around him had to exhaust themselves
overcoming barrier after barrier that stood in between the physicist and his
desire. And Hawking was well aware of the burden that he placed upon
those he knew and loved. By nature, Hawking was a fiercely independent
man.
Hawking had been suffering from a cough, and that evening, the cough
worsened. This, by itself, was not unusual; people with ALS often have
problems keeping their airways clear as their chest and throat muscles
deteriorate. Hawking had frequent coughing fits and occasional bouts of
pneumonia. But this time was different. Hawking was turning blue, and
Laflamme and the nurse tried to convince him to see a doctor. Hawking
refused. “He said, ‘I think I’ll be okay.’ I said, ‘Stephen, I think you need to
see a doctor,’” says Laflamme. “And he might have mentioned that he didn’t
want to waste anybody’s time. ‘It’s just going to take an hour. And if you’re
okay, we’re all good. If you’re not okay, then there will be somebody.’ And
fortunately, he said yes.” So at one in the morning, Laflamme—the only
member of the party who spoke French—found a doctor willing to make an
immediate house call. When the doctor saw Hawking, she was shocked.
“And then she came back to the kitchen and said, if he doesn’t go to the
hospital, he will not make the night.”
So they hustle Hawking into the car and drive wildly through the streets of
Geneva, Laflamme behind the wheel and the nurse in the back seat trying to
navigate with a map, and eventually make it to the hospital, where Hawking
is immediately admitted. And then there they waited, not knowing what was
going on. Sometime several hours later, Laflamme remembers, a doctor came
in, a bit out of breath. “She said, ‘I had to do it… to intubate him. He stopped
breathing.’” In order to get air into Hawking’s lungs, the doctors had to
anesthetize him and put a tube down his trachea. He would remain
unconscious so long as the tube was in place. It wasn’t at all clear that he
would ever regain consciousness.
According to her memoir, Jane was at a campsite in Germany and called
Geneva to check in. She had no idea anything was wrong until she reached
Hawking’s assistant: “‘Oh, Jane, thank goodness you called,’ she almost
shouted down the phone. ‘You must come quickly, Stephen is in a coma in
hospital in Geneva and we don’t know how long he’ll live,’” Jane writes.
“The news was shattering. It plummeted me into a black pit of anxiety, misery
and above all, guilt.”40
Two days later, a doctor approached Jane with a terrible choice. “The
question that he wanted to ask me was whether his staff should disconnect the
ventilator while Stephen was in a drugged state or should they try to bring
him round from the anaesthetic.” To Jane, there was no question whatsoever;
they must try to save her husband. But if Stephen survived, he would never be
the same. To allow him to breathe, they would have to give him a
tracheostomy, inserting a tube through an opening in the front of his throat,
below the larynx. To keep him from choking to death, Stephen would require
constant attention from nurses who could suction and maintain the tube, and
keep him from acquiring a life-threatening infection. And since air would no
longer be flowing past his larynx or into his mouth, he would no longer be
able to talk.41
Nearly thirty years after his tracheostomy, Hawking told the BBC that the
operation had driven him to attempt suicide. But without assistance, someone
in his condition would be hard-pressed to figure out how. He tried holding
his breath, he said. “However, the reflex to breathe was too strong.”42
The physicist was almost entirely locked in. Unable to write, unable to
talk, barely able to move, and entirely dependent on people whom he
couldn’t communicate with, in the early fall of 1985 Hawking was in a
desperate state. Jane tried to take care of the immediate financial issues; she
instructed his student Brian Whitt to work on revisions to the Brief History
manuscript. On the advice of Kip Thorne, she applied to the MacArthur
Foundation to help get funding for the new, expensive nursing care Stephen
needed. (There was little alternative because the NHS was so ill-suited to
handle a patient like Stephen outside of an institutional setting.) But even
these matters paled by comparison with Stephen’s need to figure out a means
of connecting to the world outside his head.43
At first, Hawking could communicate only by means of an alphabet card.
An assistant would point to the letters and Hawking would raise his eyebrow
to indicate which one was correct. However, this was such a slow means of
communication that it would functionally mean the end of his career; there
was no way that he could do physics or write books or give lectures at a
pace of a word every two or three minutes, using the full capacities of two
adults the whole while. Little better was a primitive buzzer system by which
the physicist could indicate one from a small number of words or commands;
this was suitable to indicate his immediate needs, such as whether he needed
his limbs to be moved or something similar, but utterly useless beyond that.
There had to be another solution.44
Half a world away, a Silicon Valley engineer, Walt Woltosz, had designed
a communications system for his mother-in-law, Lucille, who had suffered
from ALS and died in 1981. A colleague of Hawking’s heard about the
project and rang Woltosz up. “I got a call from a physicist, and he said ‘I
know you’re working on computer systems for people with ALS, and I’ve got
someone in England, he’s a professor of physics who lost the ability to
speak, and he needs a system,’” Woltosz later said in a video. The system,
dubbed the Equalizer, allowed the user to apply very simple inputs—such as
a button press or a click of a switch—to select elements displayed on a set of
menus. It was laid out so that he could quickly indicate a number of
frequently used words, or, if need be, to spell out words letter by letter. He
could store sentences for later use or recall them when required, and the
program would even attempt to reduce the number of clicks needed by
anticipating his next word choice, similar in many ways to the predictive-text
algorithms in modern cellphones. And then, on command, the program could
send the sentence to a speech synthesizer. Once Woltosz discovered that the
person in need was Stephen Hawking, he sent over the necessary equipment
—an Apple-II computer linked to a hardware speech synthesizer made by the
Speech Plus company—and Hawking took to it like a drowning man takes to
a bit of flotsam. It was nothing short of liberation for the professor. His only
complaint, which he frequently joked about, was that the speech synthesizer
had an American accent.45
The computer system was big and bulky, far too large to move around
with Hawking and his wheelchair. But, outside the Newnham Croft School,
which Timothy Hawking had started attending around the time his sister
graduated, Stephen had a lucky run-in. There was an engineer whose children
also attended the school, and Stephen asked him whether he could help adapt
the Equalizer system to fit on a motorized wheelchair. The engineer—David
Mason—could, and did, and soon became the UK distributor for Woltosz’
system. What’s more, David’s wife of ten years, Elaine, just happened to be a
nurse.46
When she met Stephen, Elaine was freshly back in the work world (too
soon for her taste) after taking a hiatus in her career as a nurse midwife to
raise her two young children. After a bit more than a year working at
Addenbrooke’s Hospital in Cambridge, the opportunity to work with Stephen
proved too tempting to pass up, and she soon became a mainstay of the
nursing rotation. (Jane was having difficulty finding nurses and keeping them
around.)47
Despite three difficult months of hospitalization, the loss of the ability to
speak, and the financial strain that new nursing requirements put on the
household, to all appearances the Hawking family was actually better off
than it had been before the tracheostomy. Jane was no longer bearing the
majority of the burden for Stephen’s care; that role fell upon the nurses. And
Stephen no longer needed a translator to be understood; paradoxically,
having lost the facility for speech, he was able to communicate more
efficiently and more quickly than he had in years, thanks to his new computer
system. The difference was the most profound to Timothy. “The first four or
five years of my life, my dad was able to speak with his natural voice, but it
was very, very difficult to understand what he was saying. For me as a three-
year-old, I had no understanding of what he was saying. So I didn’t really
have any communication with him for about the first five years of my life,”
Timothy told a BBC interviewer, Dara Ó Briain, in 2015. “It was only when
he got the speech synthesizer that I was actually able to start having
conversations with him. It was kind of ironic in a way that him losing his
voice was actually the start of him and I being able to form a relationship,
really.”48
But Stephen still took the transition hard. He was now even more
dependent on a support network of people around him. “For him it was a step
downwards. He felt he was giving in to his condition,” Jane told reporter
Brian Appleyard. “There were so many instances when a practical step for
everybody else meant, for him, giving in—an admission of defeat.”
Stephen Hawking did not admit defeat easily, even when he would play
with his children. “He used to play board games,” Timothy told Ó Briain.
“He wasn’t the easiest opponent, particularly at chess.”49
“Surely he let you win?” asked an incredulous Ó Briain.
“Well, no. There was no compassion there at all. He was hugely, hugely
competitive.”
Timothy wasn’t the only child who experienced his father’s fierce
competitive streak. A number of years after Laflamme graduated, he visited
Hawking with his two children, Patrick and Jocelyne, in tow—then about
eight and six years old. At Elaine’s suggestion, they decide to play a board
game together, one called Avalanche. Each player, in turn, drops marbles into
the top of a box full of little obstacles that swing back and forth; as the
marble falls, it might get stuck on one of the obstacles, or, if dropped in the
right place, it might jostle some of the obstacles just so and, trigger a cascade
of falling marbles. As it’s a physical game, Hawking couldn’t do it alone; he
directed Laflamme where to drop the marble. “So there is Stephen Hawking
playing against my two kids. Patrick was not very good. He loses very
quickly,” says Laflamme. “But Jocelyne was six or seven. She has a good
flair for maths and those kind of puzzles, and she’s really competing against
Stephen.” Laflamme laughs. “I had to put a marble in a place, so I was going
from one location to the other, and I’m waiting for Stephen to make a sign…
and, I don’t know, something caught him and he kind of winks, and I let it go
down. And he loses. And Stephen looks really pissed. And then he says, ‘I
didn’t! I didn’t say yes!’”50
As Laflamme well knew, Hawking’s stubborn refusal to admit defeat
extended to his scientific work, too. During Hawking’s convalescence,
Laflamme kept beavering away at the arrow of time problem, and no matter
what he did, the math was leading him in a different direction from what
Hawking’s intuition was saying. The collapse of the universe was not
behaving like a time-reversed version of its expansion; the arrow of time
would not simply switch direction as the universe begins to recollapse.
Upon Hawking’s return, Laflamme showed the results to his mentor, who
refused to believe him. “I was not getting that things were going to reverse.
And Stephen would kind of challenge me back. And he would say, ‘Well,
you’re doing this, but have you thought about this approximation?” Laflamme
says. “And of course, I hadn’t thought about it. So I would go back and
calculate for a couple of weeks, and then come back and I say, ‘Well, okay, I
think this approximation is okay, because of this, this, this, this. And I still
don’t have the arrow of time reversing.’ But he said, ‘Oh, but what about
that,’ and then he would send me off for another two weeks.” Hawking had
an endless litany of objections, each of which could be defeated eventually,
but bringing Laflamme no closer to convincing his adviser that his arrow-of-
time assumption was wrong.51
Luckily, Don Page, a Hawking student from more than a decade back,
came to town in early 1986, and sat in on one of the Hawking students’
weekly Friday meetings. Page had been working on the same problem, but
from a slightly different direction—and he was getting the same answer as
Laflamme. “Don, being older and more mature than me, says, ‘Stephen will
never believe it unless it comes from himself,’” Laflamme says. “So our job
is not to go kind of frontward, straight to his face, telling him he is wrong. We
have to ignore the fact that he’s wrong and build the case slowly, by moving
things together until he will realize the idea is wrong.” Page and Laflamme
set out to do just that. They presented him with little results that they had
proven, bit by bit, without any reference to the arrow-of-time problem, each
of which Hawking accepted. Yet though these little results were seemingly
unrelated, once you accepted them, together they excluded the possibility that
the arrow of time would reverse. “And then, suddenly, one day, Stephen says,
‘But the arrow of time thing—this idea will never work!’ Then both Don and
I said, ‘Absolutely.’”
Hawking had to make a few tweaks to his draft of Brief History, deleting
some of his earlier claims about jugs jumping off the floor and the like, and
adding a paragraph in which he publicly admitted his error:
At first, I believed that disorder would decrease when the universe
recollapsed. This was because I thought that the universe had to return
to a smooth and ordered state when it became small again. This would
mean that the contracting phase would be like the time reverse of the
expanding phase. People in the contracting phase would live their
lives backward: they would die before they were born and get younger
as the universe contracted.… However, a colleague of mine, Don
Page, of Penn State University, pointed out that the no boundary
condition did not require the contracting phase necessarily to be the
time reverse of the expanding phase. Further, one of my students,
Raymond Laflamme, found that in a slightly more complicated model,
the collapse of the universe was very different from the expansion. I
realized that I had made a mistake: the no boundary condition implied
that disorder would in fact continue to increase during the
contraction.52
Considering how close he had been to death—and what his recovery had
cost him—Hawking was functioning again in a remarkably short time. And
thanks to his new computer system, and his round-the-clock nurses,
subsidized, in part, by charitable organizations, Hawking in some ways was
functioning better than ever. Though he was still totally dependent upon
people around him, the nature of that dependency had shifted in subtle ways.
By distributing his nursing needs over a larger, more professionalized team,
he had reduced the burden on his students, his children, and especially, Jane.
Indeed, he had mostly unyoked himself from his wife. And with his computer
system, he could communicate more freely than ever before; he no longer had
to rely upon a translator.
The computerization of Hawking’s voice certainly came with its quirks
and annoyances. It didn’t slur words, which was important to him (“One’s
voice is very important,” he would say. “If you have a slurred voice, people
are likely to treat you as mentally deficient.”) But he could not inflect his
intonation to express emotion or emphasis. And because the screen was
visible not just to Hawking but to anyone looking over his shoulder, there
was no longer any privacy in the act of composing a sentence; it is as if
people could tap into his premotor cortex and divine what he was going to
say before he actually said it. A number of his friends and colleagues
describe chatting with Hawking as something very different from speaking to
anyone else. “It was participatory,” says director Errol Morris. “You could
sit next to him, and as he was writing, you could read what he was writing on
the screen. And it became a kind of guessing game. How is he going to finish
that sentence? What’s going to come next? What is he saying?” (“He almost
always wanted to complete his own sentences,” Thorne recalls. “I soon
learned not to complete his sentences for him. There was one exception: if
we were in a huge hurry and I had a taxi [en route] to take me to the airport,
then I was permitted to complete a sentence so we could move rapidly.”)53
But the voicebox and its voice quickly became so deeply associated with
Hawking that he himself couldn’t imagine replacing it even as better
synthesizers came around. Neither could his family; that robotic voice was as
much a part of Stephen as his corporeal body. As Jane put it, “it was now so
closely identified with Stephen’s personality that I found it quite upsetting if
one of the children played around with it.”54
The computerized voice—the most visible aspect of the disaggregation of
his bodily functions across a network of external people and devices—also
helped cement the physical-infirmity-begets-genius leitmotif. It raised
Hawking once and for all to the realm of transhuman, which, unfortunately,
sometimes made it harder for him to engage in the ordinary social
interactions of humanity. As he told BBC interviewer Ó Briain in 2015: “I
sometimes get very lonely because people are afraid to talk to me or don’t
wait for me to write a response.”55
But those who knew Hawking were relieved that the physicist had so
quickly recovered his strength from his near-death experience. By early
1986, Hawking had turned back to book-writing with renewed vigor. There
was a lot of work yet to do, and Guzzardi and Hawking and his students set
about revising the manuscript, kneading it and kneading it until it was ready
to go to production.v Nor did Hawking’s scientific production slacken. He
and his students were publishing at the same sort of pace in 1987 and 1988
as they had been before Geneva. On the surface, at least, everything seemed
back to some semblance of normal.
And normalcy suited Hawking just fine. Despite his disability, his deepest
desire was not to be perceived as anything abnormal, just as a particularly
gifted scientist. He didn’t want people to concentrate on his personal
struggles, or to pity him. Most of all, he didn’t want to be turned into the star
of some movie-of-the-week sob story. “Neither I nor my family would have
any self-respect left if we let ourselves be portrayed by actors,” he wrote.56
At this point in his life, for Hawking, it was all about the physics.
Footnotes
INFLATION (1977–1981)
Christmas season, 1977. Cambridge is cold and dark. Very dark. At this
time of year, the sun begins its retreat before tea-time, and has surrendered
the field by 4:30. Jane Hawking had spent the past few hours caroling—she
had recently joined the choir of St. Mark’s Church in Newnham—and, with
Lucy in tow, headed back toward her home where Stephen was waiting for
her. Alongside walked Jonathan Hellyer Jones, the choirmaster. In her
memoir, Jane later wrote, “I talked as I had not in years. I had the uncanny
sensation that I had met a familiar friend of long acquaintance.” Talk of
music, of travel, of faith. Of loss. Jonathan had lost his wife, Janet, to cancer
just a year and a half prior. As the three walked through the darkness,
illuminated periodically by the headlamps of a passing car, an intimate
friendship bloomed.1
Jonathan, an organist as well as choirmaster, began giving weekly piano
lessons to the seven-year-old Lucy. “At first he came strictly for the length of
the lesson, then he stayed a little longer to accompany me in the Schubert
songs I was learning,” Jane writes. “After a few weeks of this routine,
Jonathan began to come early for lunch or stay for supper afterwards and to
help with Stephen’s needs, relieving Robert of all the chores which had
oppressed him for so long.”2
Robert, the eldest Hawking child, was then only ten years old, and had
been put in an unusual and difficult position for a kid his age: caretaker to his
father. And Jane was rightly worried about the effect this was having on her
son, who was growing moody and withdrawn. “Much though he loved and
respected his father, it was obvious that Robert needed a male role model,
someone who would romp and tussle with him, someone who would ease
him out of a childhood already lost into adolescence, someone who would
not expect anything of him in return, least of all assistance with their own
physical requirements.” (Many years later, when asked how motor neurone
disease had diminished his life, Stephen said, “When my children were
young, I missed not being able to play with them physically.”) Jane soon
learned that Jonathan was willing to fulfill that role. “When we had got to
know him a little better, Robert would lie in wait for him by the front door
and pounce on him on his arrival, throwing him to the floor and wrestling
with him. Jonathan took this unconventional form of greeting in good part and
responded in kind to a growing boy’s need for a good rough-and-tumble to
release his excess energies.”3
In May 1978, when the pair were sitting in one of the small chapels that
ring the perimeter of Westminster Abbey, Jonathan made clear that he was
committed to Jane and to her family, “come what may.” Though the two had
not yet consummated their relationship, Jonathan had become a pillar for her,
and he was to become a part of the Hawking household.4
Quite naturally, Stephen resented the relationship and was hostile at first.
Stephen eventually acquiesced, telling her that “if there was someone who
was prepared to help me, he would not object as long as I continued to love
him.”5
When Stephen and Jane were married in 1965, they were barely able to
hope it would last thirteen months, much less thirteen years. They were just
as unaware in 1978 that this unconventional arrangement would mark the
midpoint of their marriage: the next twelve years would be the story of Jane
and Stephen and Jonathan, not just Jane and Stephen.
In his late thirties, Stephen Hawking was no longer a wunderkind. The blush
of youth had faded from his work on black holes—as revolutionary as his
insights had been, the Golden Age of Black Holes was over.
The blush was off of his marriage as well. The idealism of the first
decade and the changes in Stephen and Jane’s life together had given way to
practicalities—as much as Stephen wanted to hold fast against the progress
of his condition, day by day he had to make concessions to the people who
had to care for him, and surrender little bits and pieces of his independence
and his dignity.
Even so, Hawking adapted. He realized there was more to physics than
coming up with new theorems—and more to life.
It therefore seems that Einstein was doubly wrong when he said, “God
does not play dice.” Consideration of particle emission from black
holes would seem to suggest that God not only plays dice but also
sometimes throws them where they cannot be seen.9
(Some two decades later, by the time he was selling out the Albert Hall, the
joke had finally reached a state of high polish: “Thus it seems Einstein was
doubly wrong when he said, God does not play dice. Not only does God
definitely play dice, but He sometimes confuses us by throwing them where
they can’t be seen.”)11
Nevertheless, Hawking had reached a wide audience on his own terms.
There was not a single word in his Scientific American article about motor
neurone disease, or wheelchairs, or disability. It was all about the physics.
You get the permission, first from our internal authorities of the
theoretical group, then permission from the [Lebedev Physical]
Institute, then it was sent somewhere else, and we finally get the
permission to submit it somewhere. If it is submission for a Soviet
journal, this would not be a problem; it was very quick. But if it is
submission for anything going abroad, then it would go first to the
Academy of Sciences. Then from Academy of Sciences, it would go to
a special place where censors were checking that we did not say
anything unallowable about the beginning of the universe. All of these
would take maybe two months, three months, depending. Then it will
return to us with permission, hopefully. Usually, yes. After we get
permission, if it is something to be sent abroad, then we need to
translate it into English and somebody will translate.… If we first
make a preprint, then we actually got two separate permissions, one
for preprint and another for sending it, for example, to Physics
Letters. And it was not possible to make a Xerox copy; we must type
it twice, we must insert all equations twice. And then we send it to
Physics Letters or somewhere abroad, then usually it will travel from
us to the journal a month, two month[s] depending, and when they
reply, the replies would come to us with the same time delay.13
Much better were conferences, where scientists could meet face to face,
share ideas, and make friendships. However, the cold war was still very
much underway in 1981—Ronald Reagan, Margaret Thatcher, and Leonid
Brezhnev gave little hope for a détente anytime soon—and relatively few
Western physicists ever went to conferences in the Soviet Union, much less
forged relationships with Russian physicists.
One rare exception: Stephen Hawking.
Hawking’s Russian connection came through his longtime friend Kip
Thorne. In the mid-1960s, Thorne had attended a London lecture by a
prominent Russian physicist, Igor Novikov. “After Novikov’s lecture, I
joined the enthusiastic crowd around him and discovered, much to my
pleasure, that my Russian was slightly better than his English and that I was
needed to help with translating the discussion,” Thorne later wrote. “As the
crowd thinned, Novikov and I went off together to continue our discussion
privately.” The two struck up a friendship, and Thorne had an entrée into the
world of Russian physics. Thorne visited Moscow for several weeks in 1969
and again in 1971 as the guest of Novikov’s collaborator, Yakov
Zel’dovich.iv In 1973, he brought Hawking along with him. Since then, every
few years, Hawking made it over to Moscow. Along with Thorne, he had
become one of the main conduits for Russian cosmological ideas to flow to
Europe and the Americas.14
In the fall of 1981, Linde’s work—he had written up three papers to
explain his idea—was still bouncing around somewhere inside the Soviet
bureaucratic machine. But it so happened that there was a conference on
quantum gravity in Moscow, and Hawking was in attendance. Linde
presented his idea at the conference, and Hawking, in the audience, listened
attentively, but didn’t make any comment.
The next day, Hawking was invited to give a lecture at Moscow State
University, and Linde, who had come to hear the talk, found himself
unexpectedly drafted to act as translator. So Linde got up on stage—in front
of all the finest minds in Moscow—and began the laborious act of
translating. Hawking would say a word or two. Then, since these were the
days before Hawking’s computer, Hawking’s student would have to interpret
what Hawking was saying, and then Linde would translate the student’s
words into Russian. After half an hour of this, Hawking started discussing
Linde’s idea of slow-roll inflation: “Steve said, ‘There is this interesting
suggestion of Andrei Linde,’ and I happily translated,” Linde recalls. “And
then he said, ‘But the suggestion is completely wrong.’ And for the next half
hour, in the presence of all the best people in Moscow, I was explaining
what’s wrong about my paper.”15
After the excruciating talk, Linde told Hawking that he had been a faithful
translator, but he disagreed on a number of crucial points—and asked
whether the esteemed professor would like to discuss the matter further in
private. The two repaired to a smaller room with a blackboard and started
arguing about the details of Linde’s proposal. “Apparently, all the institute
was in a panic, because a famous British scientist had just disappeared in the
middle of Moscow,” recalls Linde. Eventually, Hawking asked Linde
whether he would like to continue the discussion at his hotel. “Oh, my God! I
did not have any permission,” Linde remembers thinking. “To hell with the
permission!” Linde followed Hawking to the hotel. The discussion turned
from physics to family—they showed each other photographs—and they
were soon fast friends. Before they parted, Hawking told Linde he was
planning a conference in Cambridge for the summer of 1982, and invited the
Russian to attend. It would be a conference to remember.16
By studying the themes and imagery in these kharjas, and in various other
medieval poetic traditions, Jane thought she could help untangle their
evolution throughout different regions of Spain and their spread to other parts
of Europe.
But Jane’s hopes came crashing into reality soon after Robert was born,
and she gave up “whatever illusions I might have held about combining
motherhood with some sort of intellectual occupation.”22
Though she could occasionally find snatches of time to work on her
thesis, Jane didn’t find a great deal of support from her husband, whose
“contempt for medieval studies was unrelenting” as she labored on and off
for more than a decade to complete it. Now, with a baby on the way, Jane
pushed to finish her dissertation—which she did with just days to spare.
Thirteen years of effort, but it paid off; she defended her thesis a year later
and then was officially a doctor of philosophy. It scarcely mattered that the
degree didn’t lead to a university teaching position, or to any further work in
medieval literature. She started teaching French to primary-school students
after-hours, and eventually wound up as a part-time tutor for high-school
students. “I won the respect of my pupils and gradually discovered a
professional identity for myself,” Jane writes. “I was awakening from an
intellectual coma.”23
She also found an outlet in monthly meetings of a new local antinuclear
group, Newnham Against the Bomb. At the same time that Stephen visited
Moscow in October 1981, the Hawkings mailed a letter to their friends and
colleagues around the world expressing “alarm at the acceleration of the
arms race.” The pair had been antinuclear activists early in their
relationship; it was one of the activities that had drawn them together, and
Jane believed that it was doing so again. “The tendency for us to slip into the
roles of master and slave was arrested. We were companions and equals
again, as we had been in our campaigning in the 1960s and early 1970s,” she
writes in her memoir. However, even as she found fulfillment in her activism,
in teaching, and eventually, by helping Jonathan plan concerts, the nagging
doubt remained that Stephen didn’t take her pursuits seriously. As Jane put it,
she had “become used to regarding [her]self as inferior and [someone]
whose endeavours were usually either met with disdain or ignored.”24
At almost the exact same moment in October 1981 that Jane sent the
antinuclear letter to everybody in the Hawkings’ Rolodex, Stephen and
fellow Cambridge physicist Gary Gibbons were sending out invitations to a
select group of physicists around the world.
The Nuffield Foundation—a charitable organization set up in the 1940s
by automobile magnate William Morris, Lord Nuffield—had agreed to
sponsor a series of scientific conferences. Hawking and Gibbons had
decided that, given all the excitement surrounding the theory of inflation, the
next one should be devoted to the “very early universe”: the first second after
creation. The workshop they planned would stretch over three weeks at the
end of June in Cambridge—and all the leading lights in the field would be
invited.
Since the news of inflation had begun to break in 1980, physicists around
the world had been playing around with the idea in hopes that they could
make the idea work; even though Guth’s model was broken, it so neatly
solved a number of key problems nagging cosmologists that it sparked an
international effort to try to shore it up.
Andrei Linde’s paper describing his “new inflation” idea emerged from
the Moscow bureaucracy in late 1981 and circulated informally as a preprint
at the same time that it was submitted to Physics Letters B, which then sent it
out for peer review. One of the reviewers was none other than Stephen
Hawking. “As a friend of Linde’s, I was rather embarrassed, however, when
I was later sent his paper by a scientific journal and asked whether it was
suitable for publication. I replied that there was this flaw,” Hawking writes
in A Brief History of Time. The problem was that like Guth’s inflation,
Linde’s new inflation would produce universes unlike our own. But Hawking
and his student Ian Moss suggested that by fixing some of the
oversimplifications of Linde’s model, they could generate a realistic
universe—not perfectly realistic (it would have too much matter), but no
longer suffering from gross flaws. As a reviewer, Hawking made an unusual
suggestion to Physics Letters B: “I recommended that the paper be published
as it was because it would take Linde several months to correct it, since
anything he sent to the West would have to be passed by Soviet censorship,
which was neither very skillful nor very quick with scientific papers.”
Linde’s paper came out in February 1982, and a Hawking/Moss paper fixing
the model up a bit was published a month later.25
In late 1981, Paul Steinhardt and his student Andreas Albrecht had
independently come up with a similar idea to Linde’s new inflation scenario;
they submitted their work to Physical Review Letters in January 1982, and it
was published in April. Hawking despised Steinhardt as a result—Hawking
had come to believe that Steinhardt and Albrecht had plagiarized Linde,
having gotten the idea from a lecture Hawking gave in Philadelphia shortly
after the Moscow conference. This is even though Steinhardt and Albrecht
referenced Linde’s work—which had since circulated in preprint—in their
Physical Review Letters paper. At the time, though, they were blissfully
unaware of Hawking’s (and Linde’s) anger, and happily toiling away at trying
to figure out what sort of universe would emerge from a new-inflation-type
model.
One important feature of any inflationary-type scenario is that the way our
present universe looks is closely linked to the roughness, the foaminess, of
spacetime on the smallest scales. The laws of quantum mechanics ensure that
very tiny objects (like the tiny proto-universe prior to inflation) aren’t
completely uniform. When inflation hits, those non-uniformities, those
fluctuations, get blown up incredibly rapidly. Places where the proto-
universe was slightly more dense would become vast swaths of space where
galaxies form in abundance; less-dense places would become voids, empty
spots where few stars or galaxies will ever be born. So, one big challenge
for cosmologists was to try to use the model to predict precisely the effect of
the inflationary blow-up on those density fluctuations—to explain exactly
how the quantum-mechanical fluctuations in the proto-universe would leave
their imprint on the universe after inflation finally turned off. And scientists
around the world were getting vastly different answers with their
calculations. Nobody could tell whether new inflation was working or not.
This was a solvable problem, and the Nuffield workshop was the ideal
place to solve it. Most of the key players were going to be attending. Alan
Guth at MIT. Mike Turner and Jim Hartle from Chicago. Paul Steinhardt at
the University of Pennsylvania. Jim Bardeen at the University of Washington.
Frank Wilczek at Santa Barbara. John Preskill from Harvard, to name a few.
And, thanks to Hawking’s Russian connections, a large group of Soviet
physicists (now including Andrei Linde) were there as well.
“There was somebody who never said anything, posing as a scientist and
wearing a black suit—that was the KGB guy—but they had a very
distinguished Russian delegation with [Alexei] Starobinsky and Linde,”
Preskill recalls. “And I remember two things about them. One, the World Cup
was going on, and they’d be glued to the TV watching the soccer game. And
the other was they would xerox papers like mad, because it was very hard to
get a xerox machine back then in the Soviet Union if you wanted copies of
scientific papers.… They were always at the xerox machine, in the xerox
room making copies.”26
At the start of the conference, nobody could agree on the proper value for
the density fluctuations. “People come and start talking with Steve and telling
him that the amplitude of the fluctuations is large,” Linde says. “But they did
not know, because there was a group of people who were saying that, yes,
they are very small. And then a group of people were saying, we don’t really
know what they look like, but they are large. And Starobinsky did not discuss
much with anybody because he has problems with his speech—stuttering.”
For three weeks, the physicists attended lectures and then broke up into
groups to calculate. Hawking, Guth, and Starobinsky worked solo while
Steinhardt and Turner teamed up with Bardeen. (Steinhardt apparently didn’t
know why Hawking was giving him the hairy eyeball.) And after a time, the
different groups—and their different approaches—revised their calculations.
(Except for Starobinsky. “He just made the statement which he made before.
He did not care,” says Linde.) The answers began to converge—Guth and
Starobinsky and Steinhardt’s team all were getting the same answer, more or
less. Hawking was a late holdout; he had published a preprint which made an
initial calculation that disagreed with the consensus by several orders of
magnitude. Toward the end of the workshop, Stephen gave a lecture, and
Guth, Steinhardt, and other physicists “were poised to challenge him when it
came to the crucial, controversial, step” in his calculations, Guth writes.
“But it never happened! Hawking’s lecture followed the preprint until the
end, but at the last step he substituted a new calculation.… True to
Hawking’s style, he did not mention that he was correcting his preprint, or
that his new answer was basically in agreement with Starobinsky or me.”27
By the end of the conference, everyone was in agreement. They had an
answer. And the answer was: new inflation didn’t work; it would create a
universe populated with black holes rather than one filled with stars and
galaxies.
It was a bittersweet ending. New inflation was dead, having lived for
barely six months. But in its failure, it pointed the way forward. With their
detailed calculations, the scientists had figured out what was going wrong—
and realized what conditions would be necessary for a new-new-inflation-
type scenario to work. By the end of the workshop, in fact, Turner and
Steinhardt had proposed just such a model; new inflation might be dead, but a
newer inflation would soon take its place. Better yet, the Nuffield
cosmologists had managed to generate some predictions about what kind of
universe one might expect after inflation—predictions that could ultimately
be tested by careful observations of the cosmic microwave background and
other astrophysical phenomena. Despite the early failures, inflation would
become a mainstay of modern cosmological thought.28
Stephen Hawking was at the very center of the inflation revolution. Even
though he himself didn’t come up with the idea, he was a key catalyst that
rallied the community around the idea, developed it, and triggered its
widespread adoption. To be sure, Hawking had done some important work
on inflation at Nuffield, and with his student Ian Moss, he had not just
tweaked Linde’s model, but incorporated inflationary ideas into his nascent
no-boundary scenario.
Hawking’s contribution to inflation was not that of a junior scientist
performing novel calculations, but that of a senior scientist spreading ideas,
helping to develop them, and connecting communities of scientists so they
could work together. At forty, Hawking was no longer a wunderkind. But
Nuffield proved that Hawking was still at the center of the cosmological
universe.
The early 1980s was an exciting time for Hawking. Even though he
suspected that he wouldn’t come up with many more big original ideas like
the no-boundary scenario, much less Hawking radiation, it scarcely mattered.
He was beginning to establish a new persona as an established, mid-career
physicist. Stephen Hawking had made a name for himself in the scientific
community, and he had amply proven that he deserved his reputation.
One downside of being a mid-career scientist is that you have to care more
about politics and relationships—you have more weight, and you’re expected
to throw it around to support your allies. And in the wake of Nuffield,
Hawking did just that to help his friend Andrei Linde. Hawking had been
muttering about Steinhardt for a few months before Nuffield, and the problem
came to a head shortly thereafter.
Michael Turner and John Barrow (who, like Hawking, was a former
student of Dennis Sciama) were writing up a summary of the workshop for
Nature. When they passed a draft along to Hawking for comment, Hawking
apparently took exception to a passage that gave Steinhardt and Albrecht co-
credit with Linde for coming up with the idea of new inflation. Hawking
reportedly suggested that the pair drop the credit for Steinhardt and Albrecht,
or, failing that, dilute their claim by giving Hawking and Moss credit as well.
They took the latter course, but told Steinhardt about the behind-the-scenes
maneuvering. After a heated back and forth with Steinhardt, Hawking backed
down, seemingly accepting that Steinhardt and Albrecht had come up with
their result independently. The matter was closed. For the moment, at least.29
Hawking also began to enjoy the platform that he was beginning to
acquire as a world-renowned expert on relativity. He was getting used to
delivering public lectures; with all the awards he was receiving, he had to
speak more and more frequently. And with more and more gravitas. Some of
his lectures, such as an address he presented in 1980 to inaugurate his tenure
as Lucasian Professor, were considered important events in their own right—
the Lucasian talk, entitled “Is the End in Sight for Theoretical Physics?,” was
reprinted in Physics Bulletin.
Hawking, ever the optimist, seemed to be growing more convinced that
the “end” of physics—the formulation of a grand unified theory—was likely
to happen before the turn of the century. Making such predictions is always
dangerous. Indeed, Hawking recalled a few cases where scientific
soothsayers had been proven wrong by history; for example, he quoted
quantum scientist Max Born as declaring that “physics, as we know it, will
be over in six months” in the late 1920s, upon learning about Dirac’s
equation.v But Hawking made predictions with all the confidence of a mid-
career scientist flushed with success. Though some of the details would
change over the years, and some of the more out-there suggestions (such as
that physicists might soon lose their jobs to computerized scientists) would
disappear entirely, Hawking’s thoughts were gaining an audience beyond the
small circle of physicists who truly understood his work.30
Like Jane, Stephen was trying to find fulfillment as his circumstances
shifted. His physical condition was slowly but surely getting worse. He had
lost the ability to write in the mid-1970s; a few years later, he was barely
able to scrawl a signature. When Hawking was elected to the Lucasian chair
in 1979, he was asked to sign a big book that all the teaching officers at
Cambridge University sign. “They brought the book to my office, and I signed
with some difficulty,” Hawking explained a number of years later. “That was
the last time I signed my name.”31 vi
In addition, despite his outward confidence, the nagging doubt remained
about whether his best days were behind him mentally as well as physically
—whether, as he told a reporter in 1982, he was over the hill with respect to
generating brilliant and novel ideas in physics. Indeed, the mathematical
precision that marked Hawking’s early work was already beginning to give
way to a vagueness that was only partially masked by the boldness of his
ideas. “I’ve given up being rigorous,” he told an interviewer in 1983. “All
I’m concerned about is being right.”32
In retrospect, there was a seed of truth in Hawking’s fears; the early
1980s marked the last time that Hawking was producing truly novel ideas in
physics. The Nuffield conference and the no-boundary proposal were the last
really significant scientific contributions of his career. While he would
continue to do good work, nothing that Hawking would produce would come
close to what he had done before. His reputation from here on would rest on
something other than his physics.
Footnotes
i There was actually a third, known as the monopole or relic problem, which
was Guth’s initial motivation, but it’s not necessary to go into that particular
detail to understand the reasons for inflation’s success.
ii With the discovery of dark energy, scientists came to realize that the
universe wasn’t really just coasting, but slowly accelerating in its expansion.
But that came two decades after Guth proposed inflationary theory.
iii The fabric of the early universe was trapped in an uncomfortably
energetic state—something akin to a supercooled fluid, which is trapped in
the liquid phase well above its freezing temperature. When the universe was
trapped in this unstable state, the fabric of spacetime expanded exponentially
quickly. But after a certain point (and the precise details of how this
transition occurs are a major part of what distinguishes different variants of
inflationary theory), there was a phase transition akin to the supercooled
fluid suddenly freezing, where the rapid expansion stops.
iv Zel’dovich, like Thorne’s PhD adviser, John Wheeler, had worked on his
country’s thermonuclear weapons project. The physics of compressing matter
at extremely high temperatures is interesting in fields other than black-hole
theory.
v Interestingly, I have not yet been able to find any credible source for this
quotation. Other literature that uses this quotation ultimately seems to trace it
back to the Lucasian lecture or A Brief History of Time. A few months before
the Lucasian lecture in 1980, Hawking wrote to Born’s son, Gustav, stating
that he had “heard” the quotation in question: “I wonder if you have any
information on the accuracy of this story or any further details that would be
relevant.” Gustav’s reply: “I am afraid that the assertion attributed to my
father is news to me.” Gustav then suggested that Hawking contact a
professor at Gottingen University. Letter from Stephen W. Hawking to Gustav
Born, February 14, 1980, Churchill College Archives, BORN 1/4/2/25;
letter from Born to Hawking, February 19, 1980, Churchill College
Archives, BORN 1/4/2/25. Unfortunately, I have found no further
correspondence on the matter or any indication of where or from whom
Hawking might have heard the quotation. “I’m not familiar with the quote,”
Born biographer Nancy Greenspan told me in a series of e-mails.” She
added, “I don’t believe Born would have said it.”
vi As fitting as this would have been, it appears not to be quite true. A 1981
book (the proceedings from a different Nuffield conference held in 1980)
bearing Hawking’s signature—with solid provenance—was sold at a
Christie’s auction in 2019. (It went for £8,750.) See “Superspace and
Supergravity Signed by Hawking,” Stephen Hawking, 1981, Lot 51,
Christie’s, https://onlineonly.christies.com/s/shoulders-giants-making-
modern-world/superspace-supergravity-signed-hawking-51/70082.
CHAPTER 12
The pair’s rivalry was good-natured, and they occasionally teamed up when
it came to physics or pranks (at one point, the two snuck a peacock into a
friend’s bedroom). But there was a bit of sharpness at times—it was a
rollicking, complex, love-hate relationship that was totally unlike anything
the Hawkings would have seen at ultraconservative Cambridge.3
Even in the world of physics, where professors were expected to be a
wee bit quirky, Feynman, in particular, stood out. He not only flouted the
norms, but he made sure everybody knew he was flouting them. For example,
Feynman spent many of his afternoons at a Pasadena strip club—five to six
days a week, by his own account—and he was proud to serve as an expert
witness when the proprietor of the club was charged under obscenity laws.
“When my calculations didn’t work out, I’d watch the girls,” Feynman
testified. The Los Angeles Times headline blared, “Bottomless [Club] Helps
Nobel Physicist with Figures.”4
Feynman’s seemingly effortless ability to attract attention apparently
grated on Gell-Mann. One Feynman friend wrote:
W hen Hawking arrived at Caltech at the end of August 1974, he was there
to work not with Feynman or Gell-Mann—who had both made their names in
particle theory—but with his general-relativist friend Kip Thorne. Thorne,
like Feynman a generation before him, had been a student of John Archibald
Wheeler before landing at Caltech. But despite the recent advances in the
field, Thorne was the only one there at the time doing general relativity. (“I
do not think that I have the independence, as Kip does, to forge relativity out
of the wilderness,” another of Wheeler’s students wrote his adviser.) So, in
addition to his general relativity work, it was inevitable that Hawking got a
great deal of exposure to particle physics and particle theory—including
Feynman’s path integral method.7
Hawking’s ability to use his hands and arms was continuing to
deteriorate; by the time he reached Caltech, he was no longer writing
equations or drawing pictures, and by the time he left a year later, he was
unable to feed himself. Feynman’s path integral method gave Hawking a tool
well suited to his very visual way of doing physics problems—and, thanks to
his prodigious memory, Hawking not just absorbed the method, but mastered
it, to the point where Feynman himself seemed to be a little touchy about it. “I
remarked to Feynman that I was impressed by Stephen Hawking’s ability to
do path integration in his head,” a Feynman friend wrote. “‘Ahh, that’s not so
great,’ Feynman replied. ‘It’s much more interesting to come up with the
technique like I did, rather than to be able to do the mechanics in your
head.’”8 A few years later, Gell-Mann couldn’t resist throwing a bit of shade
as well:
Upon his return to Cambridge in 1975, Hawking was promoted and at last
given a permanent position—a “readership,” a junior form of professorship
—and some stability. Previously, Hawking had bounced around Cambridge
with a series of fellowships and research assistantships—not by itself
unusual for a person his age, but more and more untenable given his
condition and his increasing renown. After all, Hawking had become one of
the world’s leading experts on black holes; any scientist interested in those
odd collapsed stars would encounter Hawking’s work. As a result,
Hawking’s name was likely to come up in any scientific discussion about
black holes, no matter how farfetched.21
For example, in 1973, two physicists at the University of Texas at Austin
suggested that the 1908 Tunguska event—a mysterious blast that flattened
hundreds of square kilometers of Siberian forest—had been caused by a
miniature black hole smashing through the Earth. Miniature black holes had
been dreamt up by none other than Stephen Hawking, so any story that
mentioned the Tunguska theory had to mention him. (In fact, Hawking’s
second appearance in the pages of the New York Times was in a 1974 article
about Tunguska.)22
There were maybe a dozen or two theorists—at Cambridge, at Caltech, at
Moscow, at Princeton—who knew that the recent work on black holes
heralded more than just a better understanding of a new and bizarre type of
object in the heavens. It was a key to reconciling the clash between quantum
theory and relativity. And one discovery by Hawking, in particular, promised
to blow the field wide open. As a front-page story in the New York Times put
it a few years later:
When Hawking discovered that black holes could radiate and evaporate in a
burst of energy (more on this discovery in the next chapter), it had all sorts of
unexpected consequences—consequences that Hawking and his colleagues
were just beginning to grasp.
The past decade of work, what Kip Thorne dubbed the Golden Age of
Black Holes, had proven beyond a doubt that black holes had incredibly
troublesome properties. In 1964, Roger Penrose showed that at the center of
a black hole, there had to be a singularity, a point where the fundamental
assumption that spacetime is a smooth manifold no longer holds; the laws of
relativity must break down when a black hole is born. However, that
singularity is shielded from view, entirely blocked off from the rest of the
universe. The singularity is shrouded by the event horizon. Beyond that
imaginary boundary, there is no means of escape from the black hole—not
even a means of communicating anything to an outside observer.
Shortly thereafter, Werner Israel, Brandon Carter, Stephen Hawking, and
other physicists formulated the no-hair theorems: the idea that a black hole
has almost no distinguishing features whatsoever; once you know how much
a black hole weighs, how fast it is spinning, and how much electrical charge
it carries, there’s nothing else to say. There is nothing else you can say; mass,
charge, and angular momentum are the only information that someone outside
of the event horizon can extract from a black hole. Everything else about the
black hole is lost, inaccessible to anyone unwilling to cross the event horizon
after it. That’s what general relativity says, anyhow.
Quantum mechanics also has something to say—its laws dictate absolute
limits on extracting information from a system. It’s impossible to know both a
particle’s position and its momentum with perfect accuracy; after a certain
point, getting information about one means losing information about the other.
Just as important, although less well explored by scientists at the time, was
the implication that information cannot be totally annihilated—it can be
moved from place to place, stored, transformed, gathered, scattered—but it
can’t be totally erased. This principle is hardwired in the mathematical
structures that describe quantum objects, thanks to a property known as
unitarity. Without this property, the quantum mechanics doesn’t really make
sense; the logic of quantum interactions is no longer perfectly consistent.
So quantum mechanics says that information can’t be lost under any
circumstances, while general relativity says that when information falls into a
black hole, it can’t escape. These two statements aren’t necessarily in
contradiction; perhaps the black hole is storing the information, keeping it out
of sight forever behind the event horizon. That was perfectly acceptable to
both the relativists and the quantum theorists. Everything was copacetic.
Until Hawking messed it all up.
Hawking’s 1974 discovery that black holes radiated energy was
counterintuitive and troubling. And he soon discovered that it had a very
disturbing implication. When a black hole radiated energy, the energy had to
come from somewhere; in fact, it came from the mass of the black hole itself.
So as a black hole radiates, it gets less massive—smaller—over time. As the
black hole gets smaller, Hawking’s calculations implied that its radiation
becomes hotter and brighter, making the black hole shrink even faster.
Smaller. Brighter. Smaller. Brighter. The process can take billions of years,
but eventually, the black hole, in theory, shrinks to a point and explodes in a
flash. Black holes aren’t permanent; they explode.
It took many of his colleagues a few years to realize how important this
revelation was. Hawking—perhaps because of his exposure to particle
physics at Caltech—quickly understood that the impermanence of black holes
drove a major wedge between the rules of general relativity and quantum
theory.
Quantum mechanics says that information can’t be lost under any
circumstances. Relativity says that when information falls into a black hole,
it can’t escape. And now, Hawking argued, black holes must eventually
themselves be destroyed: they can’t store information indefinitely behind an
event horizon. Someday, far in the future, the event horizon must finally be
ripped away, leaving nothing behind. So where did the information go? It
couldn’t have been stored; it couldn’t have escaped past the event horizon; it
couldn’t have been encoded in the Hawking radiation (for the nature of
Hawking radiation implies that it couldn’t carry information).
That’s really the essence of the black-hole information paradox. Quantum
theory says that information cannot be destroyed. Yet Hawking’s description
of a black hole implies not just that information is lost to an outside observer
during a black hole’s lifetime, but that a black hole’s lifetime is finite—and
no information survives its demise. It’s a dilemma that sits squarely at the
intersection of quantum theory and relativity. Two theories giving
contradictory answers—both can’t be correct. Perhaps, just perhaps, by
figuring out what’s going wrong in one or the other analysis (or both),
Hawking could take a large step forward in reconciling these two
mathematical frameworks. By delving into the heart of a black hole, Stephen
Hawking might have pointed the way to a new master theory: one that would
surpass quantum theory and relativity just as those theories had surpassed
Newtonian mechanics.
The first step, though, would be to convince others that there was a
contradiction. Toward the end of his stay at Caltech, Hawking wrote up a
paper outlining the information theory paradox and submitted it to Physical
Review D. It was published more than a year later—likely the sign of a battle
among peer reviewers who took issues with Hawking’s argument. And
though the work was noticed (and cited) over the next few years by a handful
of relativity theorists, particularly those in Dennis Sciama’s and John
Wheeler’s circles, the black-hole information paradox took a relatively long
time to diffuse into the wider scientific community. Even Leonard Susskind,
who would spend much of his career wrestling with the implications of the
paradox, didn’t hear about it until 1981—fully half a decade later.
As Susskind tells it, he first heard about the black-hole information
paradox at a quirky little gathering in the attic of a very controversial
millionaire: Werner Erhard. Erhard was the leader of a self-improvement
movement that grew rapidly, gained high-profile celebrity followers like
John Denver and Yoko Ono, and was soon likened by some to a cult.
Erhard pumped some of the money from those self-improvement trainings,
known as “est,” into a foundation—and with that foundation, he managed to
attract some of the world’s leading physicists around him. Starting in the
mid-1970s, Harvard’s Sid Coleman, Caltech’s Dick Feynman, and a number
of other leading physicistsii had been periodically wined and dined and
entertained at a physics conference paid for by the largesse of the est
Foundation. Coleman, who helped recruit attendees, had no illusions about
how participants might feel about the “self-improvement courses” that
bankrolled Erhard’s fortune:
Hawking didn’t let anyone outside his intimate circle see even a hint of his
struggle; he would wheel through the streets of Cambridge day after day to
work, and was proud of how seldom he failed to show up. He labored for
hours together with his students or by himself, completely absorbed by the
physics problems swirling inside his head.33
There were plenty of problems to work on. His time at Caltech had left
him with greater interest in particle physics and the techniques peculiar to
that field, especially the Feynman path integral method. Starting in 1976,
along with Jim Hartle and Gary Gibbons, he spent several years exploring
and refining the ideas that would eventually lead to his no-boundary proposal
in 1981. Then there was the black-hole information paradox; he had
convinced himself that information had to be lost when it fell into a black
hole, but he hadn’t yet come up with an explanation for where it might go.
Hawking would think about this on and off for years, eventually coming to the
conclusion that the particles falling into a black hole “go off into a little baby
universe of their own”—that the information wasn’t truly lost, but was
carried off into a region that bubbled off our own sheet of spacetime, forming
its own tiny cosmos. The black hole, the ultimate engine of destruction, was
also a source of creation. (Of course, Hawking was forced to abandon this
notion when he conceded that information wasn’t lost in a black hole.) So
Hawking was not showing any signs of running out of ideas.34
It was those ideas—that brain—that made people take notice of Hawking,
and allowed him to climb the academic hierarchy so quickly. Alas, the perks
of such a high academic position weren’t financial; even at its peak,
Hawking’s Cambridge salary was never enough to sustain him in a
comfortable manner, especially given his medical expenses. The chief benefit
was respect. And Hawking was getting plenty of that. Sometimes that respect
even translated into a cash dividend.
The Einstein Award, for example, was a sign of respect that came with
$15,000 and a gold medal. It was a somewhat controversial honor that had
been given out, off and on, since the 1950s. Controversial because it had
been put together by former US Atomic Energy Commission chairman Lewis
S. Strauss, who was reviled by much of the scientific community for
engineering the downfall of J. Robert Oppenheimer, the scientist who, during
World War II, had led America’s project to design the atom bomb.With few
exceptions, the award had gone to bomb designers, mostly hawkish ones.
(When presented with the award in 1954, Richard Feynman didn’t want to
accept it until physicist Isidor Rabi—an outspoken dove—convinced him to.
“You should never turn a man’s generosity as a sword against him,” Rabi
reportedly said.)35
In August 1977, Strauss’ son sent out a letter to former winners of the
prize, including Feynman and John Archibald Wheeler, asking for advice.iii
“Do you think that Hawking’s work is of sufficient importance to justify the
Einstein Award, quite apart from the humanitarian aspects of the matter?”36
Wheeler’s response left little doubt. He explained that he had co-written a
textbook on gravitation in the early 1970s that had included biographical
sketches of the important scientists in the field of gravitation, from Galileo
onward. Only three of those sketches were of living people: Bob Dicke of
Princeton University, Roger Penrose, and Stephen Hawking.iv “I have no
question he belongs among the immortals,” Wheeler wrote. (Feynman’s
response was less effusive. His whole response to Strauss’ letter reads: “In
reply to your letter of August 2, I would most certainly agree that Dr. Stephen
W. Hawking’s work is deserving of the Einstein award.”)37
The following January, Hawking was awarded the prize. “The work being
done by Hawking and his group may lead to a Unified Field Theory which
has been the Holy Grail of 20th century physicists,” the award announcement
declared. “It was sought, unsuccessfully, by Einstein for the last 40 years of
his life.”38
Like Einstein, Hawking would never set eyes on the grail. But it’s not
necessary to complete a quest to achieve immortality. And as Wheeler had
predicted, Hawking was already being counted among the immortals of
science.
In late 1979, officials at Cambridge decided that Hawking would become
the next Lucasian Professor of Mathematics. The Lucasian Professorship was
created in the 1600s by order of Charles II to fund, in perpetuity, a “man of
good character and reputable life… soundly learned and especially skilled in
mathematics.” The second person to hold the position was none other than
Isaac Newton—and from then on, the Lucasian Professorship was the most
prestigious academic position in the world. Its holders were instantly
considered (rightly or wrongly) to be the intellectual heirs to Newton
himself. Many of the subsequent Lucasian Professors—Charles Babbage,
Gabriel Stokes, Paul Dirac—were legendary figures in mathematics and
physics, adding to the already nearly mystical prestige of the office. And
now, Hawking was counted among them. Nobody could any longer question
Hawking’s position as a first-rate scientist, or doubt that his growing fame
was based upon his important contributions to theoretical physics.39
Nobody, that is, except Hawking himself.
“I think I was appointed as a stopgap to fill the chair as someone whose
work would not disgrace the standards expected of the Lucasian chair, but I
think they thought I wouldn’t live very long, and then they could choose
again, by which time they could find a more suitable candidate,” Hawking
told an interviewer a number of years later. “Well, I’m sorry to disappoint
the electors.”40
Footnotes
i The nature and amount of radiation emitted by the black hole depends on
how small it is, so, in theory, a very small black hole—much smaller than
anything known to exist—might be visible as it emits light or, eventually,
explodes.
ii The list of people attending one 1977 gathering reads like a who’s who of
cutting-edge particle physics. It includes several future Nobel winners:
Curtis Callan, Geoffrey Chew, Sid Coleman, Ludwig Fadeev, Dick Feynman,
Jeffrey Goldstone, Yoichiro Nambu, David Gross, Roman Jackiw, T. D. Lee,
Larry Susskind, and Gerard ’t Hooft, to name a few.
iii Wheeler, despite being a hawk on the Lewis Strauss and Edward Teller
side of the Oppenheimer schism, never suffered as a result. He was almost
universally beloved by the physics community.
iv When Hawking sent Wheeler some biographical information in 1972, it
may have been the first professional use of his famous “born-300-years-
after-Galileo died” line. Letter from Stephen Hawking to John Archibald
Wheeler, June 12, 1972, Wheeler Papers, Box 11.
PART III
INSPIRAL
The years from 1970 through 1974 were the peak of Stephen Hawking’s
scientific life. Two major results, two big insights about black holes, were
enough to establish him as a physicist of the first rank.
Hawking was laboring in an increasingly crowded field. There were
several other physicists, from the United States, from Russia, from England,
and from Israel, who were all working together—and against each other—in
divining bizarre and unexpected properties of black holes. But it was
Hawking who finally came to the key realization, one so strange that even he
didn’t believe it at first. It was Hawking whose name would be forever
linked with the understanding that black holes weren’t really black. And this
discovery became the crowning achievement—and the coda—to the Golden
Age of Black Holes.
The soul of a black hole is its singularity. The idea of a point of infinite
curvature where the laws of physics no longer make sense—that’s the puzzle
that first drew physicists into studying black holes. They wanted to
understand the pathological heart of these collapsed stars. And in the late
1960s and early 1970s, nobody understood singularities like Roger Penrose
and Stephen Hawking. Hawking had made his name on singularities. But by
1970, Hawking was beginning to turn his attention not to the singularity, but
to another part of black-hole anatomy: the event horizon.
An event horizon isn’t a physical thing any more than the equator of the
Earth is a real line; it’s an abstract mathematical concept. Luckily, physicists
were more or less agreed on the definition of that concept. An event horizon
was a boundary around the black hole that marked the point where even light
wasn’t moving fast enough to avoid crashing into the singularity. That is, the
event horizon was an imaginary surface where the escape velocity—the
speed an object needed to be moving to get out of the black hole’s clutches—
was light speed. This was the natural, obvious definition. But it had subtle
problems. For example, as a star collapsed into a black hole, the full-sized
event horizon would suddenly pop into existence once the surface of the star
got small enough. Instantaneous, discontinuous changes lead to problematic
math.
In 1970, Hawking realized that a subtle change in the definition of the
event horizon could eliminate those problems. Instead of thinking of the event
horizon as a place where the escape velocity was equal to the speed of light,
he thought of it as a boundary between two separate realms: the realm where
objects were doomed to hit the singularity versus the one where objects
could still belong to the outside universe. Hawking’s definition, which he
dubbed the absolute horizon, wasn’t quite the same as the old escape-
velocity definition, now termed the apparent horizon. It takes a bit of playing
around with Penrose diagrams to see the differences, but they’re there. For
example, when a star collapses into a black hole, an absolute horizon doesn’t
suddenly pop into existence fully grown like an apparent horizon does; it
starts off as a point in the middle of the star and grows and grows as the star
collapses, eventually bursting outward beyond the surface of the shrinking
star.ii
It’s a somewhat confusing and very subtle difference that doesn’t mean a
heck of a lot to a nonspecialist. But to Hawking, who was painting Penrose
diagrams of collapsing stars in his head, it was a terribly important
distinction that forced him to look at black holes in a new way; it changed his
mental picture of the geometry of black holes. And one night in late 1970, not
long after his daughter, Lucy, was born, he was exploring his new mental map
while getting ready for bed. All of a sudden, he was struck by a geometric
insight.
Hawking was picturing how light rays behaved when on the absolute
event horizon, and his new definition made it clear to him that those rays
could never approach each other. They could move parallel to each other, or
get farther apart, but they couldn’t get closer together, because if they did, at
least one of them would be captured by the black hole. This is a really tough
concept that Hawking struggled to try to explain to non-scientists. In A Brief
History of Time, he tried to give his readers a sense of this geometric insight
by likening it to robbers trying to escape from cops: “It would be like
meeting someone else running away from the police in the opposite direction
—you would both be caught!”2 iii
That geometric insight led to another. The behavior of light rays on the
event horizon determines the shape that the event horizon can take, not just
through space, but through time. Saying that light rays on the event horizon
can’t approach each other is equivalent to saying that the horizon can never
get smaller over time; it must always get bigger, no matter what. And when
two small black holes merge to make one big black hole, the event horizon of
the big black hole must be at least as big as the combined areas of the smaller
black holes. That is, the geometric rules of general relativity ensured that
black holes could never shrink; they had to grow, or at the very least stay the
same size. It was a new law of nature, a novel insight not just into the
behavior of black holes, but also into the nature of spacetime itself. This was
what became known as Hawking’s area theorem.
“I was so excited with my discovery that I did not get much sleep that
night,” Hawking later wrote. “The next day I rang up Roger Penrose. He
agreed with me.” By Hawking’s telling, Penrose had, in fact, been thinking
along the same lines, but hadn’t yet grasped its significance.3
“The thing about the area theorem, this needs to be straightened out,
because I don’t know,” says Penrose. “You see, I’d been in Cambridge to
give a talk, and after the talk, I went to Stephen’s office, and we talked about
various things. And I talked about the area of a black hole in general, and the
fact that the area had to increase over time.” The two discussed the idea for a
while, and puzzled a bit about the nature of event horizons, because, in
general, their shape could be quite complicated and might defy analysis.
Penrose then went back to his hotel room and went to bed. “Early the next
morning, Stephen phoned in, and he said he got a new idea. And he said, ‘It’s
really your idea.’”4
Nearly half a century on, Penrose finds the story of Hawking’s evening
revelation a bit confusing. The idea that black-hole areas had to increase was
already well in hand before Hawking put his daughter to bed. “Exactly what
his epiphany was, I don’t know. I thought it was that the areas of black holes
increase, but that wouldn’t be completely consistent because I have a clear
memory that he said it was your—he’s talking to me—idea,” Penrose says. “I
suspect it may have been a misunderstanding.… He might have thought I was
talking about apparent horizons.… I certainly wasn’t talking about that.”
Penrose credits Hawking with first thinking about the combined area of two
black holes colliding and combining into one, rather than merely looking at
the horizon of an individual black hole increasing over time—but that
distinction isn’t really the key insight of the area theorem, nor does it fit in
with the story of Hawking’s bedtime epiphany. “I don’t know what he
thought. Maybe he thought I had the idea but didn’t quite have it. It’s not
clear. I don’t know what the story was, really,” Penrose says. “I never
wanted to bring it up. Because it was a big thing for him.”
It was a big thing. The area theorem was a breakthrough; it allowed
theorists to constrain the behavior of black holes in ways they never could
before. For example, by applying the area theorem to the merger of two black
holes—precisely the sort of event that the LIGO gravity-wave detector
started seeing forty-five years later—Hawking was able to calculate the
maximal amount of gravitational radiation that gets released in the merger.
Too much, and it would cause the black hole to lose mass and the horizon to
shrink, violating the area theorem.
It was a profound discovery. But even Hawking didn’t yet realize just
how important it would turn out to be.
Across the Atlantic, the inspiration came from a cup of tea. Two, actually.
At Princeton, John Archibald Wheeler was pondering what happens when
you bring two cups of tea into contact—one hot and one cold. You don’t need
an advanced degree in physics to know that heat spontaneously flows from
the hot one to the cold one until they both reach the same temperature. But to
a physicist, the question of why nature behaves that way, why heat always
flows from the hot cup to the cold one and never in reverse, is the gateway to
some fundamental insights into the rules that govern our universe: rules about
how energy behaves and how time flows. It all has to do with a concept
known as entropy.
Imagine that you’ve got one hundred marbles—fifty red and fifty blue.
Pick one at random and toss it, plunk, into a shoebox a few feet away. Then
do the same to another marble. And another. Plunk, plunk. Continue until
you’ve tossed all the marbles into the shoebox. Go over and take a look;
what do you see?
You’ll (obviously!) see a jumble of red and blue marbles all mixed
together. When you dump those marbles randomly into the shoebox, they
wind up randomly arranged in the box—there’s no particular order to how
they fell, and no particular order to how they lie. Indeed, if you looked in the
box and all the red marbles were sitting on the left side and the blue marbles
on the right, you’d be astonished. The world doesn’t work that way: if you
toss marbles randomly into a box, they won’t spontaneously self-segregate by
color. Of course, it’s not impossible to have a box with red and blue marbles
neatly separated into each corner, but even in such a case, it probably won’t
last long. As soon as someone picks up the box, the marbles will jostle and
mix with each other once again. The nice neat order imposed from without
will quickly dissipate into chaos once more.
That’s entropy in action. Entropy, in some sense, is the amount of
“disorder” in a system. A shoebox with all the red marbles on the left and all
the blue ones on the right has low entropy; when someone moves the shoebox
around and the marbles mix, the entropy increases and increases until the
marbles are randomly distributed once again. It’s possible to restore order to
the system if you want; you can reduce the entropy in the shoebox by sorting
the reds from the blues. However, doing that requires you to spend energy—
and when you stop spending that energy to sort the marbles, the system’s
entropy will increase once again.
Entropy isn’t just about colored marbles in shoeboxes; the same principle
applies to all matter and energy everywhere in the universe. Red marbles
might represent “hot” molecules in a cup of tea, and the blue marbles
represent “cold” ones. Or the red marbles could stand for nitrogen in a room
full of air, and blue ones for oxygen. It doesn’t really matter. In each case, the
system will tend toward maximum entropy—total, random mixing of red and
blue—unless you expend energy to reduce the entropy. And if you don’t have
a way of pouring energy into the system to sort things out, the entropy must
inexorably increase until it reaches its maximum.
This principle is one of the most important scientific discoveries of the
nineteenth century, and it forms one of the pillars of the field of physics
known as thermodynamics: the study of energy, of heat, of temperature, of
work. The tomb of Ludwig Boltzmann, who first expressed entropy in terms
of the different configurations of marbles/atoms/molecules/objects in a
system, is inscribed with his formula for entropy:
S = k log W
Around the time of Lucy’s birth in 1970, Stephen resigned himself to using a
wheelchair; he had been able to walk short distances, albeit unsteadily, with
crutches, but that was becoming too difficult and too dangerous to maintain.
He no doubt knew that sitting in a chair constantly would hasten the atrophy
of his muscles and render him increasingly reliant on others to get about. But
he tried to get exercise in other ways. He would insist on getting himself
ready for bed—apart from untying his shoelaces and undoing his buttons, he
was able to undress himself and get into pajamas. And he insisted on
climbing the stairs to the bedroom, too, even after his legs were no longer
able to support him.7
“The way he got up the stairs was, he grabbed hold of the pillars that
support the bannister and pulled himself up the stairs with the strength of his
own arms, dragging himself up from the ground floor on up to the second
story in a long effort,” Kip Thorne told director Errol Morris a number of
years later. “Jane explained that this was an important part of his physical
therapy to maintain his coordination and strength as long as possible. At first
it was sort of heartrending to watch what appeared to be the agony of pulling
himself up the stairs until I understood it was part of life. Pulling himself up
the stairs like that.” (“There I give credit to Jane,” says Roger Penrose, who
also saw the struggle with the staircase on more than one occasion. “I can
imagine that many people would have insisted on helping him, but she felt
that this was absolutely necessary.”)8
The Hawking household was scraping by on sheer determination, but
barely. Stephen’s salary wasn’t great, and it wasn’t perfectly secure, as he
wasn’t tenured or on the tenure track. It was a stable fellowship with a
renewable six-year term, yet it was clear that Stephen was destined for
something bigger. The twenty-eight-year-old physicist was in the middle of a
burst of white-hot creativity that filled him with energy.
Just a week after Lucy’s birth, Stephen submitted a paper to the Monthly
Notices of the Royal Astronomical Society suggesting that the universe might
be filled with microscopic black holes.
Ordinary black holes—the ones that astrophysicists had been thinking
about for decades—are born when large stars, several times the mass of our
sun, collapse. Smaller, sun-sized stars simply don’t have enough
gravitational oomph to collapse down to a singularity; these become white
dwarfs or neutron stars rather than black holes. So black holes formed in this
way have a minimum mass, a bit more than twice our sun’s mass. But
Hawking realized that in the hot, dense early universe, it might be possible to
create black holes that have much smaller masses, perhaps merely as heavy
as a mountain. That sounds large, but a black hole of that size would be
roughly as small as a proton, and might even work its way into the guts of an
ordinary star and sit in its core, slowly eating it from the inside.
It was a really interesting idea, and theorists liked it for a few reasons—
applied judiciously, mini black holes could account for miscellaneous
unexplained phenomena that had been troubling astrophysicists.v
Unfortunately, Hawking based his theory on a number of faulty assumptions.
And part of his motivation was to explain some gravitational waves that an
experimentalist, Joe Weber, was detecting with his instruments. Those
detections turned out to be phantoms; his experiments were discredited a few
years later. Even so, the idea was so compelling that Hawking’s early
appearances in the popular press were, as often as not, related to his mini
black holes.
Hawking was also thinking about the no-hair theorems—he was helping
to solidify the mathematics underneath the hood. At the same time, he was
working with George Ellis, a friend and fellow former student of Dennis
Sciama’s, to finish up a treatise on spacetime, relativity, and black holes—a
book that would become a classic in the field. He even won the Gravity
Research Foundation’s essay prize in 1971, which gave him a bit of cash in
pocket.
That prize was something of a running joke in the relativity theory
community: it had been set up by eccentric millionaire Robert Babson (best
known for founding Babson College), who had a lifelong grudge against
gravity, as he held it responsible for having drowned his sister. Despite the
prize’s cranky origins, impoverished physicists entered the contest year after
year in hopes of winning the fairly generous purse. (Hawking entered half a
dozen times over the years, but only won it once. When congratulated by post
on winning such a prestigious prize, Hawking reportedly wrote back, “I don’t
know about the prestige, but the money’s very welcome.”)9
It was the most productive time in Hawking’s life. Even without the area
theorem, it would have been notable. However, the magnum opus was just
over the horizon.
Jacob Bekenstein had been hard at work for several months trying to
understand what was going wrong with the scenario posed by his adviser.
Wheeler was right: if black holes truly erased all evidence of what they
swallowed, it was irrelevant whether the infalling matter had high entropy or
low entropy. The ultimate outcome was the same. And if that were the case, a
black hole could serve as a mechanism for getting rid of entropy, for
avoiding the thermodynamic consequences of your actions. You could mix hot
tea and cold tea together without increasing the entropy of the universe, in
violation of the second law of thermodynamics.
The second law of thermodynamics says that in a closed system—one
with no energy flowing in—entropy must, without exception, increase, or at
least stay the same. And in 1971, when Hawking published his area theorem,
Bekenstein was struck by how it paralleled the second law: the size of a
black-hole event horizon must, without exception, increase, or at least stay
the same. Could there be some connection? Could the area of a black hole’s
event horizon be related to entropy? And could that harbor the answer to
Wheeler’s teacup mystery?
There were other reasons to think that the similarities between black
holes and thermodynamical systems were important. For one, another student
of Wheeler’s, Demetrios Christodoulou, had worked out some rules
regarding spinning black holes that were parallel to well-known principles
in thermodynamics.vi There were just too many coincidences on too many
fronts for it to be an accident; the rules which dictate how black holes behave
simply had to be intimately related to the rules of thermodynamics. So
Bekenstein did some calculations and confirmed that, indeed, the area of a
black hole’s event horizon behaved very much like a thermodynamical
entropy. Indeed, by May 1972, Bekenstein realized that it made sense to
define a black hole’s entropy as:
Sbh = η (kc3/Għ) A
where A is the area of the black hole’s event horizon (and k is Boltzmann’s
constant, c is the speed of light, G is the universal gravitation constant, ħ is
Planck’s constant [a value that occurs and reoccurs in quantum mechanics],
and η was an as-yet-undetermined number). This equation formalized the
tight relationship between the event horizon’s area and the entropy of a black
hole—and since a black hole had entropy, it would no longer serve as a
bottomless pit where one’s thermodynamic “crimes” could be dumped, as
Wheeler had suggested.10
Take a hot cup of tea and a cold cup of tea and pour them into a black
hole. The area of the event horizon will increase by a certain amount. Do it
again, but mix the hot and cold tea first, increasing the entropy of the system.
This time, the area of the black hole will increase more than it would if you
dumped the hot and cold tea separately. That is, the increased entropy of the
teacup system isn’t lost when you throw it into a black hole; it’s preserved as
an extra increase in the event horizon area. As Bekenstein told Wheeler,
“You have not avoided the entropy increase. You’ve just put the entropy
increase in a new place. The black hole itself has entropy.”11
Bekenstein’s formula worked beautifully; it closed the apparent loophole
in the second law of thermodynamics that was so troubling Wheeler. But
there were disturbing consequences to assuming that a black hole could have
entropy. From a thermodynamic point of view, an object with entropy had to
have a temperature. But what could it possibly mean for a black hole to have
a temperature? If you put it next to something “cooler,” would heat flow out
of the black hole? Could a black hole cool down somehow? Worse yet,
anything that has a temperature must radiate energy; after all, quantum
mechanics was born when Max Planck came up with the correct formula for
what kinds of light had to shine from an object at a given temperature. But a
black hole can’t emit light! It swallows it—that’s the very definition of a
black hole.
Even so, even knowing the seeming contradiction that his formula created,
Bekenstein decided to publish it in the physics journal Nuovo Cimento. The
article came out in August 1972, just as he was heading to Les Houches in the
French Alps for a summer school on black holes. He was eager to discuss his
ideas, as many of the world’s leading experts on black holes would be there.
Surely they would be as excited about black-hole thermodynamics as he was.
Bekenstein was still convinced that he was correct. But he couldn’t get
more traction for his ideas; he couldn’t figure out a way to resolve the
objections regarding black-hole temperature, or even figure out an exact
value for that troublesome constant η, which was still hanging out in his
entropy formula. But nobody else was working on the idea of black-hole
thermodynamics, even its opponents.
Having launched his salvo against Bekenstein, Hawking moved on to
other things. He did more work on mini black holes, and briefly explored
how the cosmic background radiation would be affected if the universe were
rotating while it was expanding. He was still plugging away at his textbook
with George Ellis—a long project that was finally coming to its end; it
would be published by Cambridge University Press in 1973. Contained in
that work—and in a paper that came out a wee bit earlier—Hawking
presented a mathematical proof that (given a few assumptions) a black hole
had to be symmetric about the axis about which it spins. Hawking’s so-called
rigidity theorem put a bow around the famous no-hair theorems for black
holes—it ensured that there were no exceptions as a result of bizarre
distributions of matter in a collapsing star. As a consequence, after Werner
Israel and Brandon Carter, who were the main people behind the no-hair
theorems, Hawking and mathematical physicist David Robinson are also
considered co-creators.
That September, Stephen accompanied Kip Thorne to Warsaw and then to
Moscow for the first time since his student days. Jane overcame her fear of
flying and joined her husband overseas. It was a glorious visit; his suite in
the Hotel Rossiya overlooked the parti-colored onion spires of St. Basil’s
Cathedral, and Thorne, who spoke fluent Russian, showed them the town,
including Tchaikovsky ballets and Borodin operas at the Bolshoi. The only
substantial complaint was the suspicion that the hotel was infested with bugs
—of the electronic variety.16
On that trip, Hawking met Alexei Starobinsky for the first time—taciturn
because of his stutter, but happy to share what he was working on. As it
happened, Starobinsky and his mentor, Yakov Zel’dovich, had a piece of
research that Hawking found fascinating.
Hawking well knew that it was possible to extract energy from a spinning
black hole; indeed, the no-hair theorems were based, in part, upon how
spinning black holes have to emit gravity waves under certain circumstances,
carrying energy away from the black hole. But Starobinsky and Zel’dovich
realized that a very rapidly spinning black hole could lose energy not just by
radiating gravity waves, but also by radiating particles of all varieties—
photons, electrons, neutrinos, and the like.
It was an intricate argument, but the idea was inherently quantum
mechanical. As described in Chapter 5, the uncertainty about energy and time
built into quantum mechanics ensures that on the tiniest scales and for the
shortest times, particles are constantly winking in and out of existence. These
vacuum fluctuations are the reason why the quantum theoretic picture of
space is frothy and constantly churning, rather than smooth and continuous as
relativity would have it. These quantum fluctuations aren’t figments of
mathematical imagination; under certain circumstances, such as if you pour
enough energy into a small enough space, these evanescent, now-they-exist-
now-they-don’t particles are given enough oomph to fly away and be
detected by a particle counter.ix Starobinsky and Zel’dovich argued that in the
vicinity of a fast-spinning black hole, the violent twisting of spacetime in the
region around the event horizon would do just this, and send particles
shooting outward in all directions.
Hawking was intrigued by the Muscovites’ argument; he tended to believe
it, unlike Kip Thorne, who had already made a friendly wager with
Zel’dovich that the calculations were wrong.x But he was unconvinced that
the pair had done the math quite right. They were venturing into almost
uncharted territory. The important work on black holes thus far was pretty
much classical physics. Gravitational radiation and the like could be handled
purely with the equations of general relativity without trying to tangle with
the clashing principles of quantum theory. But once you started mucking about
with particles, you needed to take quantum behavior into account. Classical
physics wouldn’t cut it, but because relativity and quantum theory clash, it
wasn’t easy to make the two work together. So, while Hawking believed the
result, he didn’t think that Zel’dovich and Starobinsky had done the
calculations in a sufficiently rigorous way. “I didn’t like the way they
derived their result,” Hawking said in 1984, “so I set out to do it properly.”17
Hawking spent weeks absorbed in thought, blasting Wagner operas at high
volume on the record player, and ignoring all distraction. As Jane later
wrote, he was “transported to another dimension, lost to me and to the
children playing around him.” He often wouldn’t respond to Jane’s attempts
to elicit a response out of him, so absorbed was he in his mathematics. Jane
soon grew to hate Wagner.18
Stephen’s physics took a “semiclassical” approach: he chose some
elements of classical general relativity and added quantum theory when
necessary to describe the behavior of matter. Almost by definition,
semiclassical gravity is wrong on some level; it’s a mishmash of sometimes
internally contradictory theory, but with luck, it’s good enough to yield some
interesting insights. And so Hawking spent the late fall trying to figure out
whether spinning black holes would emit particles. By the end of November,
he had done enough work to get an answer: yes. There was radiation. But he
hadn’t yet figured out how much.
Lots, it turned out. Too much. Way too much. “I was expecting to discover
just the radiation that Zel’dovich and Starobinsky had predicted from rotating
black holes,” Hawking wrote in A Brief History of Time. “However, when I
did the calculation, I found, to my surprise and annoyance, that even non-
rotating black holes should apparently create and emit particles at a steady
rate.” Something was wrong.19
Hawking’s calculations showed that even when a black hole wasn’t
rotating, the mere existence of an event horizon would affect the vacuum
fluctuations nearby; the black hole would swallow some of those frothy,
evanescent particles, while others would be liberated, freed from the
confines of the black hole’s gravity, and zoom off into the cosmos. A distant
observer would see those liberated particles as radiation—it was as if the
event horizon were emitting particles in all directions. Even though the black
hole itself would never allow a particle to escape, it could shine brightly
nonetheless. As Hawking wrote in Brief History, “particles do not come
from within the black hole, but from the ‘empty’ space just outside the black
hole’s event horizon!”20
Hawking checked and rechecked his calculations in late December and
early January, convinced that he had made an error somewhere, that perhaps
the imperfect half-quantum, half-classical mathematical approach was letting
him down. He simply didn’t believe that a motionless black hole could
radiate particles, even though his calculations were coming to a remarkable
conclusion: the particles streaming away from the black hole had precisely
the characteristics predicted by Max Planck in 1900 at the dawn of the
quantum revolution.
As described in Chapter 5, Planck was trying to figure out the properties
of the light that was emitted by an object at a given temperature. To get the
right answer, he had to make an uncomfortable assumption—that light came
in discrete packets known as quanta. But after Planck did so, his formula did
a beautiful job; it predicted exactly the sort of radiation that a nonreflective
object, known as a black body, would emit at any given temperature.
Planck’s derivation of the so-called black-body spectrum was the first
triumph of quantum theory. Whenever it appeared in nature, it was a sign of
the quantum-mechanical underpinnings of the universe—it had profound
connections to the ultimate workings of nature.
Hawking’s calculations were showing that the radiation coming off of a
black hole had a black-body spectrum. More, the black-body spectrum was
precisely that predicted if the black hole had a “temperature” defined by the
black hole’s surface gravity—just as Bekenstein had predicted!
Hawking, Bardeen, and Carter had objected to Bekenstein’s
“thermodynamics” arguments because black holes didn’t have temperature—
they didn’t have temperature because they couldn’t radiate. Now Hawking’s
calculations seemed to show not only that black holes do radiate, but that
they radiate just like any other object with a fixed temperature does.
Bekenstein had been right all along. At first, Hawking didn’t want to tell
anybody about his results, for fear that Bekenstein would find out—that
Hawking’s error would give his young rival ammunition “to support his ideas
about the entropy of black holes, which I still did not like.” Consequently, in
early January 1974, he told only his closest friends about the puzzling result
he was getting.21
Hawking was having a hard time believing what his calculations were
telling him, but Martin Rees immediately saw their significance. Dennis
Sciama was heading toward his office in early January when he ran into a
pale and trembling Rees. “He was shaking with excitement,” Sciama later
recalled. “And he said, ‘Have you heard? Have you heard what Stephen has
discovered? Everything is different. Everything is changed!’”22
Recent developments in this field have been associated with the name
of Hawking, although there have been many other significant
contributors. However, Hawking has acknowledged in his papers, and
everyone close to the subject knows, that it was Bekenstein who
initiated the field. He emphasized on physical grounds that the surface
area of a black hole is not only analogous to entropy; it is entropy; and
the surface gravity of a black hole is not only analogous to
temperature; it is temperature.32
Sbh = (kc3/4Għ) A
T = ħc3/8πGMk
Footnotes
i At the time, the French term for black holes was astres occlus (occluded
stars); the French were still resisting the obscene-sounding name trous noirs
(black holes) that Wheeler had coined—made even worse by Bekenstein’s
observation that “black holes have no hair.”
ii Another way to think of it: the old definition of event horizon was, roughly
speaking, the boundary between light that can escape and light that can’t
escape at a given moment in time. Hawking’s new absolute horizon is the
boundary between light that could escape and light that couldn’t escape not
just at one point in time, but throughout all of time. As a star collapses, light
in the center of the star might be unable to escape the black hole that will
form in the future, but doesn’t yet exist, and thus doesn’t have an old-style
event horizon!
iii A slightly different way of looking at it is to remember that a particle of
light on the event horizon is just on the edge of getting caught—it has to be
traveling as fast as it can (at light speed, naturally!) on the shortest path away
from the singularity just to avoid getting captured. If two photons, A and B,
are frantically trying to escape and cross each other, then it means there are
two “shortest” paths away from the intersection: the path that A takes after
meeting B, and the path that B takes after meeting A. But there can only be
one shortest path—which means that the situation is impossible.
iv Entropy is also closely tied to the concept of information. A fuller
explanation of this relationship between entropy and information can be
found in my book Decoding the Universe.
v For example, Hawking suggested that it could be used to explain a lack of a
certain type of particle, known as a neutrino, created during fusion reactions
in the sun. The deficit was later proved to be due to neutrinos having mass,
rather than mini black holes at the core of the sun swallowing them up.
vi Specifically, Christodoulou was looking at pulling energy out of a spinning
black hole by means of gravitational radiation; he discovered that the most
efficient techniques for doing so were precisely the ones that were
reversible. This is analogous to the idea of a Carnot cycle in traditional
thermodynamics.
vii Bekenstein’s guess for η, as the equation is written in this chapter, was
(1/2 ln 2)/4π. As Hawking later showed, the true value is 1/4.
viii It’s easiest to think of surface gravity in terms of mass (the larger the
mass, the smaller the surface gravity, broadly speaking).
ix Pouring enough energy into a small enough space is precisely what “atom
smashers” are for. When scientists at the massive LHC machine at CERN see
a spray of particles in their collider, they’re really seeing particles created
out of the vacuum. You don’t even need to have the atoms smash into each
other to create those particles; if they pass by close enough to each other,
there’s still a spray of particles even though the original atoms are intact.
x This was, in fact, Thorne’s first big scientific wager. He had bet Zel’dovich
a bottle of White Horse scotch against one of Georgian cognac that a spinning
black hole wouldn’t radiate particles. He lost.
xi A number of sources say that Hawking was the youngest member to ever
have been inducted. However, there are quite a number of counterexamples.
Isaac Newton, for example, was named a Fellow in 1672 at the age of
twenty-nine, Robert Hooke in 1663 at the age of twenty-eight, and Arthur
Cayley in 1852 at thirty-one, to name a few.
CHAPTER 14
Long before humans set foot on North America, glaciers crept south from
what is now Canada. These mountains of ice gouged out great chunks of soil
and rock as they advanced. New York was almost entirely covered with a
thick sheet of glacial ice until, a mere ten thousand years ago, the climate
began to warm and the ice made its final retreat. The scars gashed out by the
glaciers filled with water and became deep and narrow lakes hemmed in by
high walls. It was here, surrounded by the post-glacial beauty of upstate New
York, that Jane and Stephen Hawking spent the second half of their
honeymoon.
It was the summer of 1965. Stephen’s PhD thesis was almost finished; he
had done most of the intellectual heavy lifting, and all that remained was to
put it into final form and submit it. Better yet, he had also secured a research
fellowship at Cambridge that would begin in the fall. It didn’t pay well, but it
came with some degree of stability, providing a decent foundation for a
newly married couple. Their honeymoon had started with a week in Suffolk
—and then, their marriage just a few days old, Jane and Stephen had bundled
into an airplane and crossed the Atlantic. A car was waiting at the airport
and drove the couple to their lodgings amid the gorges and waterfalls of
Ithaca, New York. They had arrived at Cornell University, where Stephen
was to participate in a summer school devoted to general relativity.1
“That was a mistake,” he later wrote, rather laconically. “We stayed in a
dormitory that was full of couples with noisy and small children, and it put a
strain on our marriage.” The screaming babies were a problem, but so was
the fact that they had little money for food, and wound up dining on whatever
inexpensive ingredients they could figure out how to prepare in a cheap
saucepan over a hotplate. After helping her husband—who could walk, albeit
slowly, with a cane—navigate the mile from the dorm to the lecture hall, Jane
would find a typewriter so she could type out a draft of Stephen’s PhD thesis.
As Stephen reveled in connecting with new colleagues, Jane looked with
trepidation at what the scientists’ wives had become: “They were already, to
all intents and purposes, widows—physics widows.” The Hawkings’
honeymoon was over.2
To Hawking, the summer school was definitely a net positive, even if it did
put a strain on his marriage. “In other respects, however, the summer school
was very useful for me because I met many of the leading people in the
field,” he recalled. And there were a few friends from back home as well;
there was Brandon Carter, another of Dennis Sciama’s students a couple of
years behind Stephen, and there was Roger Penrose, with his wife and sons
in tow.
For Hawking, Penrose’s arrival at the summer school was most welcome.
About a decade older than Stephen, Penrose had already established himself
as a faculty member at Birkbeck College (part of the University of London).
And he was taking an interest in the twenty-three-year-old Hawking, much as
Hawking’s adviser Dennis Sciama—who knew Penrose’s brother—had
taken an interest in Penrose a number of years earlier. In fact, it was Sciama
who got Penrose into physics in the first place. “Although I was doing pure
mathematics, he was trying to convert me to do cosmology,” recalls
Penrose.3
As a pure mathematician, Penrose had a different set of tools than
physicists did, tools that turned out to be very powerful when used on
problems in general relativity. Einstein’s equations of general relativity are
used to understand the geometry of spacetime; they describe the fabric of
space and time and how they curve under the influence of matter and energy.
But they say nothing about the overall shape, the topology, of the universe—
any more than saying that an object made out of flannel will tell you whether
it’s a shirt or a dress or a bathrobe or a tablecloth or a ball, for that matter.
The mathematical description of the fabric itself doesn’t give you a picture of
the overall shape that that fabric takes. Penrose started using topological
tools that hadn’t been used by the relativity community before, and they
happened to be extremely effective in places where spacetime gets unruly.
Like near singularities.
In late 1964, Penrose used topological arguments to prove that a
collapsing star of sufficient size would always wind up creating a singularity.
It didn’t matter how lumpy or asymmetric or irregular the star was as it
collapsed; the end result was always the same. A singularity was
unavoidable.
Penrose’s discovery caused a huge splash in the general relativity
community; he had proven something that had eluded physicists for years. For
there had long been doubts whether the singularity at the heart of a black hole
was an accident of the oversimplified physics of a perfectly symmetrical
implosion, an artifact of trying to describe messy real-world systems with the
idealized language of mathematics. Penrose had answered that question once
and for all: the monstrosity at the heart of a black hole was no figment of the
imagination. It had to be real.i
At that time, Hawking was struggling with his PhD thesis. He had come
up with a bunch of minor results—poking a hole in an alternative theory of
gravitation, exploring the formation of galaxies in the steady-state model of
the universe—but nothing really particularly notable or important. Finding a
good idea for a PhD thesis is the hardest part of the entire endeavor; it has to
be a problem that’s solvable by a relatively inexperienced researcher, yet
interesting enough to be worthy of solution. And Hawking hadn’t found that
sweet spot—until Penrose announced his black-hole singularity theorem in
early 1965.
Hawking wasn’t at the lecture where it was announced, but Brandon
Carter was, and Carter told Sciama and Hawking and George Ellis and all
the rest of the Cambridge crew about the result. Sciama asked Penrose to
come to Cambridge and present the results there. “That’s when I first met
[Hawking],” Penrose remembers. “We went into a side room where I
described more details to Stephen and George Ellis.” And then Hawking
suddenly made a connection that nobody else had. As Sciama recalled in
1990: “I remember Stephen Hawking, who was then approaching his third
year as a research student, saying, ‘what very interesting results. I wonder if
they could be adapted to understanding the origin of the universe?’”4
Penrose’s topological techniques had revealed the presence of a
singularity during the collapse of a star. Hawking realized that the same
techniques could be applied to the Big Bang model of the universe once he
made one tweak: he needed to reverse the flow of time. A Big Bang cosmos
begins from something very, very tiny and expands rapidly into something
enormous. Played in reverse, the physics looks almost the same as an
enormous star shrinking into something very, very tiny. And just as Penrose
proved that a black-hole collapse must end in a singularity, Hawking saw that
the same mathematics implies that the expansion must begin in one—with all
the difficult implications that conclusion entails.
“I realized that if I reversed the direction of time so that the collapse
became an expansion, I could prove that the universe had a beginning,”
Hawking later told director Errol Morris. “But, my proof, based on
Einstein’s general theory of relativity, also showed that we could not
understand how the universe began because it showed that all scientific
theories, including general relativity itself, break down at the beginning of
the universe.”5
It was an important and disturbing insight—and would make a superlative
PhD thesis. “It was a nice idea, and I was struck by his picking up the idea
very quickly, and being able to use it in a very general way, in a general
context,” Penrose says. However, from Penrose’s point of view, Hawking’s
real insight was not merely in reversing the direction of time—something
Penrose describes as “obvious”—but in a very technical mathematical point:
Hawking took a key element of Penrose’s proof and figured out how to apply
it not to the region near a black hole, where Penrose did, but way, way out at
cosmological distances.6 ii
Back then, Hawking was indeed as much mathematician as physicist. (“I
was trying to be as rigorous as a pure mathematician,” Hawking wrote
toward the end of his life. “Nowadays, I’m concerned about being right
rather than righteous.”) Hawking, who had a natural talent for mathematics,
picked up the techniques from Penrose quite quickly. Penrose, for his part,
was thrilled to have Hawking use his approach. He was no longer the lone
topologist in a field of geometers.7
Over the next few years, Penrose and Hawking would collaborate on a
number of projects. The most important was refining and extending the
singularity theorems. Sometimes working separately and sometimes
discussing matters by telephone, they chipped away at the underlying
assumptions of the theorems and drew more powerful conclusions. Despite
Hawking’s prodigious mathematical talent, he made some significant
mistakes. On several occasions, Hawking lofted a new mathematical proof in
a preprint or even in a published paper that wound up being wrong, only to
be corrected by other physicists. However, the errors were never severe
enough to invalidate the overarching—and correct—insights he’d had. “I call
these mistakes of the first kind: that some mistakes which are correctable and
they don’t affect the general argument,” Penrose says.8
Penrose and Hawking also wrote separate but complementary papers that,
in 1966, won each of them a version of the prestigious Adams Prize—
awarded by Cambridge for distinguished research in mathematics.
(Technically, Penrose had won the prize, but Hawking was—unusually—
given a supplementary award.) The pair’s collaboration culminated in 1970,
when they combined their ideas about singularities into a more general result
now known as the Penrose-Hawking singularity theorem.
But back in July 1965, during that summer school session at Cornell,
Hawking’s relationship with Penrose was just beginning—and Penrose’s
presence was no doubt appreciated as Hawking put the finishing touches on
his thesis. When Hawking turned it in that October, it was his singularity
theorem that carried the day. The first three chapters were respectable, if not
terribly innovative, but the fourth—the proof that an expanding universe must
begin in a singularity—that was exceptional work. It was that theorem that
got Hawking his PhD.9
It was also what began Hawking’s life-long honeymoon with physics.
W hen the Hawkings got back to England, the newly married couple
immediately had to tangle with the unpleasant practicalities of setting up a
household with very little money in their pockets. Stephen had luckily landed
a research fellowship with one of the residential colleges at Cambridge—
Gonville and Caius—which would give the two just enough money to live on
for the time being. But it didn’t provide them with any housing, and the pair
struggled to find a place to live. Not that they could really settle in even if
they had found a domicile; Jane still had to finish up her undergraduate
degree in London, so she would take the train from Cambridge on Monday
morning, spend the week there, and then hustle back to Stephen on Friday
afternoon.10
Stephen could still get along on his own reasonably well; he could walk
to and from work with his cane, and though his speech was deteriorating, he
could make himself understood. But there was always the knowledge that the
disease would be tightening its grip. It was already happening.
During the summer school at Cornell, Stephen was seized by an
uncontrollable paroxysm of choking, a coughing fit that left him rumpled and
exhausted—the first such fit Jane had witnessed. Even though she knew her
husband was battling a terminal illness, her helplessness in the face of the
onslaught was a shock. Hawking’s father, Frank, was a physician—actually,
one of the world’s most renowned experts in tropical medicine—and though
it was a bit out of the area of his expertise, he put his son on a regimen of
oral and injectable B vitamins. Sadly, there was little else he could do for his
son.11
Despite the shadow cast by the disease, Jane and Stephen began to carve
out a life for themselves, with a little bit of help from their parents here and
there. And Stephen was beginning to make a reputation for himself as a
physicist. His thesis launched his career in style. It resulted in three peer-
reviewed papers in quick succession, the first of which was published in
Physics Letters the day after his wedding. His pride in the accomplishment
was probably only slightly diminished by the fact that the primary author was
one “S. Hawkins.” The Hawking singularity theorem was winging its way
around the world, and people began to take notice of the young physicist.12
Hawking was definitely noticeable, even to people who hadn’t met him in
person. He continued to extend his work on singularities, and published quite
a few papers on that subject, as well as some of his other research on the
early universe. By the end of 1966—at the tender age of twenty-four—he had
nine publications under his belt, including one in Nature, the most prestigious
scientific journal in the world. Even his father, Frank, didn’t get his first
Nature article until his late thirties, shortly after Stephen was born. Though
Frank had wanted the young Stephen to be a doctor (and had tried to steer the
boy away from mathematics), he must have been proud of his son’s work,
even if it had nothing to do with medicine. And like Frank, who traveled all
over the world for his research, Stephen spent a great deal of time away from
home, networking with the world’s experts in relativity at every meeting he
could convince his college to send him to. At first, it was mostly the United
States—Florida and Texas were his first two trips abroad as a postdoc—but
he would soon be visiting more exotic locales.13
Now, for the first time, people were making pilgrimages to visit him.
Princeton’s John Archibald Wheeler, like Cambridge’s Dennis Sciama,
encouraged his students to visit other research groups to forge friendships
and collaborations that would help shape their careers; Wheeler scrounged
funding to allow his young charges to imbibe the “vitamins” that come with
exposure to a new group of scientists. In January 1967, for example, Wheeler
used his own money—honoraria for lectures he had given—to send a
promising young grad student, Bob Geroch, to Cambridge, so that he could
“review his work and discuss recent developments for a few days with
Stephen Hawking, who, along with Roger Penrose, is tops in the field of
singularities.”14
Writing from England in February, Geroch expressed gratitude to Wheeler
for the opportunity. “Dr. Hawking has been very nice—both in sacrificing his
time to talk with me, and in finding lodgings, food, entertainment, etc. I can
now appreciate your remark that no volume of letters can substitute for one
personal visit in producing a good relationship.” As Geroch knew firsthand,
it’s not just because face-to-face meetings create stronger bonds than letters
ever could, but also because letters can be misunderstood.15
Half a year earlier, in June 1966, a furious Hawking had sent a telegram
to Sam Goudsmit, the editor of Physical Review Letters, in an attempt to
block Geroch from publishing one of his papers. Hawking had misread one
of Geroch’s letters—he initially thought Geroch was attempting to publish
some of Hawking’s work as well as his own—but the letter merely suggested
that the two publish in tandem. Since Geroch’s work depended on
Hawking’s, Geroch didn’t want to go to press before Hawking did. After
sending out a “telegram of protest” or two, Hawking reread the letter and
realized that he had misunderstood Geroch’s intent. Hawking didn’t quite
apologize for the misunderstanding, but he did smooth things over with
Goudsmit, writing up a little summary of his work so that the two could
publish in the same issue of the journal. But, clearly, by the time Geroch
visited Cambridge in 1967, there was no ill will. Geroch imbibed plenty of
“vitamins” from Hawking and from Brandon Carter, who brought him up to
speed on the latest developments in the field.16
Travel was the lifeblood of Stephen’s early career, but travel was
expensive, and his salary was already stretched to the limit. Especially since,
in the early fall of 1966, Jane found out that she was expecting a baby.
Stephen was losing control over the muscles in his fingers and was having
difficulty writing. Yet the young couple couldn’t afford physiotherapy to try
to stave off the inevitable loss of function. Dennis Sciama intervened with the
powers that be at Cambridge to get the university to pay for twice-weekly
visits from a physiotherapist, which helped a bit. But it was clear that most
of Stephen’s writing—and all of his typing—would have to be done by
others. As he was too junior to have an assistant, that burden would fall upon
the increasingly encumbered Jane.17
Luckily for Jane, late 1966 and early 1967 marked a lull in Stephen’s
research; he had parceled off all the publishable work from his thesis into
papers already, and was looking for another line of research to tackle. As he
mentioned to Bob Geroch in the winter of 1967, research into singularities
was “asymptotically approaching a fixed number of results”—in other
words, the big questions that could be tackled by the Penrose-Hawking
approach were pretty much in hand; it was really a matter of tying up loose
ends with the singularity theorems rather than making big new breakthroughs.
It was nigh time to pull up stakes and move on to more fertile soil. It would
be three years before he made the crucial shift in his investigations from
black-hole singularities to their event horizons, leading to his next great
insight. Before that, though, Hawking would have to wander the wilderness
in search of the next big idea. For a young researcher on a contract—even
one without a baby on the way and engaged in a battle to slow down the
effects of a degenerative disease—it’s a perilous and nerve-racking time in
one’s career.18
Even in the absence of a next big idea, though, Hawking wasn’t spinning
his wheels so long as he could travel and make connections. He applied in
February to go to a small physics meeting, a “Rencontres,” in Seattle. Many
of the leading lights in physics would be there: Roger Penrose, Richard
Feynman, Charlie Misner, Tullio Regge, Bryce and Cecile DeWitt. And it
was run by John Archibald Wheeler. So Jane typed up the application (in
duplicate) and Stephen gathered the necessary recommendation from a more
senior faculty member; then he sent it overseas, fully expecting to spend the
summer in Seattle—presumably with an infant in tow.
In March, Hawking’s plans were dashed. His younger colleague Brandon
Carter was invited to the Seattle Rencontres—but Hawking was not. The
number of spots available were limited, and Wheeler had chosen Carter over
Hawking.
Carter had asked Penrose for a recommendation, but Hawking had not,
leaving Penrose in a somewhat awkward situation. Penrose didn’t want to
undermine Hawking’s chances of acceptance while helping Carter’s. So in
the recommendation, he sang Hawking’s praises as well:
The Golden Age of Black Holes was well underway, and as Wheeler was
aware, Brandon Carter was playing a major role. Hawking wasn’t—at least
not yet. He was still awaiting a second great insight, something that would
push his career forward beyond what he had done with his PhD. As
important as it was, the singularity theorem alone wouldn’t sustain his
reputation for long. And back in 1967, his bedtime flash of inspiration about
the area theorem was still three years in the future.
W ith a baby months away, Hawking had other things on his mind than black
holes. Luckily, Jane got her bachelor’s degree and no longer had to commute
to London, but Stephen was increasingly dependent upon her—from the daily
walk to and from the department to the necessaries of maintaining a
household. On top of that Jane had, somewhat optimistically, secured an
adviser for her PhD at the University of London, Alan Deyermond. The
thought was that she would write her thesis while raising a baby. And in
early 1967, she was rushing to put the finishing touches on her first research
article.
Jane’s paper had to do with a fifteenth-century Spanish tragicomedy, La
Celestina, in which an elderly procuress’ scheming leads to a shocking
number of characters dying after jumping out of windows or falling from
great heights. The central argument of Jane’s work was to lay the cause of the
tragedy to a frustrated servant’s unfulfilled longing for a mother figure. “The
idea was a fascinating one which won Alan Deyermond’s amazed approval:
he was even more amazed when I confessed that the idea was Stephen’s,”
Jane wrote. Stephen had idly thumbed through La Celestina while waiting
for his wife to finish her final examinations, and mentioned the idea to her on
the drive back to Cambridge. All Jane had to do was “to flesh out the
argument and justify the Freudian concept when applied to a text dating from
1499.”24 iii
Jane thought Stephen’s contribution to the Celestina paper was a sign of
the “harmony” in their relationship, the mutual respect for and contribution to
each other’s intellectual pursuits. It didn’t seem to occur to her that the
respect might have been one-sided. Stephen didn’t hide his scorn for
medieval studies, likening it to “collecting pebbles on the beach,” and the
ease with which he had come up with a publication-worthy idea probably
reinforced that scorn rather than lessening it.25
Regardless, the question was moot, at least for the time being. The baby’s
arrival in May put Jane’s intellectual ambitions on hold. Barely two months
after Robert’s birth, the three packed up for the Seattle workshop. “That,
again, was a mistake,” Hawking later admitted. “I was not able to help much
with the baby because of my increasing disability, and Jane had to cope
largely on her own and got very tired.” It was a long four months, and though
Stephen had earned some money in speaking fees, he hadn’t yet figured out
his next line of research. Jane, for her part, had begun to develop a fear of
flying.26
Over the summer, though, there was an important development in black-
hole research: the object that everybody had been studying finally had a
name. John Wheeler was at a meeting in New York where he was discussing
black holes with the audience—but there was no official name for these
objects. Unlike white dwarfs or red giants or neutron stars, there were only
generic-sounding names for them—“completely collapsed objects” was
common. “Well, after you get through saying ‘completely collapsed objects’
six times, you look for a shortcut,” Wheeler later recalled. “And that’s when
I found myself using the phrase ‘black hole.’”27
“It was a stroke of genius: The name ensured that black holes entered the
mythology of science fiction,” Hawking told an audience at Berkeley in
1988. “It stimulated scientific research by providing a definite name for
something that previously had not had a satisfactory title. The importance in
science of a good name should not be underestimated.”28
It was turning out to be a banner year for black holes. Just a few months
before Wheeler came up with the term for black hole, a South African
physicist, Werner Israel, had taken the first big step toward what are now
called (in another Wheeler-catalyzed coinage) the no-hair theorems.
To understand a black hole’s behavior—what it looks like to a nearby
particle, how it can move, how it can interact with the environment around it
—physicists need to know how a black hole shapes the fabric of space and
time nearby. And that’s easier said than done. Despite thinking about
spacetime around black holes for decades, by the early 1960s physicists had
only really gotten a handle on the shape of spacetime around the very
simplest possible black-hole system—the neighborhood around totally
motionless, perfectly spherical black holes. But though they could solve the
Einstein field equations for that idealized situation, they feared that actual,
astrophysical, black holes would be quite a bit more complicated. They
might rotate. They might be asymmetric and have hills and valleys, or regions
of greater or lesser density. They might pulsate or change shape. And it was
one thing to solve equations regarding idealized, simple Platonic objects;
doing the same thing with complex, messy astrophysical black holes, rather
than mere models, was incredibly difficult. It was only in 1963 that an
Australian scientist, Roy Kerr, had come up with a solution to the Einstein
equations to describe the behavior of spacetime around a spinning black
hole. The Kerr black hole was still an oversimplified object, as it was
symmetric and uniform, but it was at least one step more realistic than
before. His discovery marked the beginning of the Golden Age of Black
Holes—a rapid series of discoveries about these collapsed stars that turned
them, step by step, from a cartoonish theoretical idea into a real
astrophysical object worthy of study.
Within two years of Kerr’s discovery, Ezra Newman, a physicist at the
University of Pittsburgh, was able to take Kerr’s twirling black hole and add
one more layer of complexity: charge. Now, scientists could start examining
spinning black holes with or without electric charge, and begin deducing
their properties. Brandon Carter at Cambridge, for example, quickly became
an expert in the Kerr-Newman solutions to the Einstein field equations, and
he soon created a stir by finding completely unexpected new physics in the
way objects in orbit around a Kerr-Newman black hole move about.
So, by the mid-1960s, scientists could handle black holes that were
spinning, and that carried an electric charge. But that was still an
oversimplification. What about asymmetries? What if the black hole were
lumpy rather than perfectly spherical? In 1967, Werner Israel proved that it
didn’t matter how asymmetric or inhomogeneous a collapsing star was. The
end result was always the same: a perfectly symmetrical, spherical black
hole. No matter how bizarrely the matter might have been distributed, the
final black hole would always be an unblemished sphere. Over the next few
years, Carter and a number of other scientists extended Israel’s proof, putting
it on more and more solid theoretical ground and exploring its implications.
They figured out, for example, that any asymmetries or pulsations in the
collapsing star would wind up being radiated away as gravitational waves—
only once the black hole was as quiescent and featureless as a cue ball
would it stop radiating any asymmetries away. Even though it might be born
in complexity, a black hole always winds up extremely symmetric and
simple. It has no hair.
The no-hair theorems were great news for physicists; no longer did they
have to worry about solving equations for messy, asymmetric, pulsating
matter. When it came to black holes, that asymmetry, that mess, was all lost—
what had been a complicated lump of material had become simpler as it
collapsed. Indeed, these theorems implied that a black hole was the simplest
macroscopic object in the universe: there’s nothing more to learn—nothing
more that you can learn—about a black hole other than its mass, rate of spin,
and charge. A black hole is an almost featureless object.
It truly was a Golden Age. By the late 1960s, physicists had a whole suite
of powerful new tools: the Kerr-Newman solution, the no-hair theorems, and
Penrose’s black-hole singularity theorem. However, despite being “tops” in
singularities, Hawking wasn’t really spending much time thinking about black
holes. At the time, with a few exceptions, his work was on cosmology, on
trying to understand the conditions (and the singularity) at the beginning of the
universe, rather than astrophysical objects like black holes (and their
singularities). So Hawking had not yet played a major role in the Golden
Age. That, of course, would change within a few years.
However, in 1967, Hawking was occupied with other matters. Around the
time Robert was born, he ceased being able to walk to the department on his
own. Typically, he would hitch a ride with George Ellis, who would bring
him back and forth from work, or he would ride in a three-wheeled battery-
powered car designed for people with disabilities, occasionally bringing a
passenger along for the ride. (“I found this rather scary because I thought he
drove faster than was safe,” wrote one of Hawking’s early students.)
Hawking managed to get his research fellowship at Gonville and Caius
renewed for another two years, but the rules prevented another renewal. And
as he struggled to find another idea as powerful as his singularity theorem,
Hawking was no doubt acutely aware that some of his younger colleagues
were more than giving him a run for his money, research-wise. Not only was
Brandon Carter making a name for himself with black holes, but Martin Rees,
also a student of Dennis Sciama’s, was turning out to be a powerhouse; he
was coming up on a dozen publications in Nature on a variety of topics,
including X-ray astronomy, radio astronomy, and cosmic rays. (Rees was
skyrocketing up the academic ladder; in 1972, he would land a
professorship, and in 1973, he would be named to the Plumian chair in
astrophysics—like the Lucasian chair, a centuries-old position of high
honor.) When 1969 rolled around and Hawking’s contract expired once
more, Dennis Sciama helped wangle a “special category” of research
fellowship for his young protégé.29
At the time, there was a little-used (in fact, perhaps never-before used)
provision in the university statutes allowing for six-year fellowships for
“persons of exceptional distinction in science, literature, or art.” It probably
wasn’t intended for someone so young and inexperienced, but it allowed the
university to throw Hawking a lifeline. According to the terms of the
fellowship, Hawking would spend half of his time at the Department of
Applied Mathematics and Physics, and half at the newly created Institute for
Theoretical Astronomy, run by Fred Hoyle. Simon Mitton, the institute’s
departmental secretary at the time, recalls afternoons filled with mundane
duties tending to Hawking’s needs. “They included getting him out of his
three-wheel invalid car and putting him in the wheelchair which we kept at
the Institute of Astronomy. I used to have to give him medication, you know, a
couple of pills, and I remember having to get fixed up for his office, when he
first joined, a pushbutton telephone system in which something like twenty
lines were preloaded. It’s hard to imagine this today, but this was a massive
procedure, involving telephone engineers of what was back then a
nationalized industry having to come along and fix all of this stuff up by
hardwiring it.”30
Cambridge was doing its best to look after Hawking. The position they
gave him was a renewable six-year appointment that took the pressure off.
And, as internal documents made clear, the university had committed to
“accept[ing] the obligation of wholly maintaining him as long as he can do
his scientific work.” Nevertheless, at the time, Hawking was no closer to
winning a professorship than he had been in 1966. He needed another big
idea, and time was not on his side.31
Before Hawking’s attention turned squarely to black holes, he even flirted
with becoming an experimentalist. In 1969, a set of experiments by
University of Maryland physicist Joseph Weber was the talk of the relativity
community. Using a set of refrigerated metal cylinders, Weber claimed to
have been detecting gravity waves; the idea was that the distinctive stretch-
and-squash distortion of a passing gravity wave would set these specially
sized cylinders ringing like a bell. And Weber was claiming to have seen
dozens of gravity waves, many of which seemed to be coming from the center
of the Milky Way. It sounded almost too good to be true.
It was too good to be true. Within a few years, other physicists
demolished Weber’s claims; he hadn’t been careful enough with his
measurements and had convinced himself that he was seeing gravitational
radiation when it was mere noise. But in 1969 and 1970, Weber’s results
were being taken very seriously indeed. Hawking flew out to see Weber’s
equipment and was soon at work with one of his graduate students thinking
about how to improve Weber’s results. By November 1970, the two had
come up with a design that would be more sensitive by a factor of ten—or
even more—and Hawking applied for a grant to build their own gravity-
wave detectors.32
Luckily, the pair eventually decided not to pursue the idea. “That was a
narrow escape! My increasing disability would have made me hopeless as an
experimenter,” Hawking later wrote. “One is often only part of a large team,
doing an experiment that takes years. On the other hand, a theorist can have
an idea in a single afternoon, or, in my case, while getting into bed, and write
a paper on one’s own or with one or two colleagues to make one’s name.”33
When the next big idea came in 1970—the area theorem—and then the
huge one after that in 1974—Hawking radiation—Hawking would indeed
make his name. And he would do it alone.
In September 1969, Jane, pregnant with her second child, awoke to the sight
of her first, Robert, covered with sticky pink liquid. With a start, she realized
that the toddler had used a chair to get at the collection of medicine bottles on
the top of the refrigerator, and had drunk the lot down. Among the bottles was
a “stimulant” that Jane had been prescribed to “pep [her] up.” She
immediately ran with Robert to the doctor, who sent them off to the hospital.
As Robert began convulsing, the nurses held the little boy down and began
pumping his stomach.34
Jane watched helplessly as her child, strapped down on the cot to avoid
hurting himself, stopped flailing and sank into a coma. After some hours, his
condition stabilized. “His state was not hopeful; all that could be said was
that it was not deteriorating further,” Jane wrote in her memoir. “Coming to
my senses at this slightest of changes, I was shocked to remember that I had
left Stephen alone in the house, scarcely able to look after himself.” She had
to make a terrible decision: stay by the side of her comatose baby, or rush
home to make sure that Stephen was all right. She ran home to check on her
husband.35
Stephen was fine; he hadn’t been alone. George Ellis had come in the
morning and gotten Stephen to work. Jane returned to the hospital, not
knowing whether her son was dead or alive. When the nurse told her that
Robert had emerged from the coma, Jane could do nothing more than weep.
And while the story had a happy ending, the deep scar from having to choose
between caring for her son and caring for her husband never fully healed:
“That day,” she later wrote, “Robert survived, but a little piece of me
died.”36
Footnotes
i This insight would win Penrose the Nobel Prize in 2020, more than half a
century later; the committee described his proof as perhaps “the most
important contribution to the general theory of relativity since Einstein.”
ii The element is what’s known as the trapped surface condition. For the
proof to work properly, Penrose had to invent a precise way of describing a
curve in spacetime that enclosed the singularity of a black hole into the
future. Hawking figured out how to adapt that condition and apply it to the
universe as a whole, which was not obvious. “It’s just the observation that
you have an anti-trapped surface if you go far enough away in some standard
open cosmology,” Penrose says.
iii In fact, Jane’s justification for the Freudian concept occupies a paragraph
at the beginning of the fourteen-page document and presents neither a deep
nor a rigorous argument. Jane Hawking, “Madre Celestina,” Annali—Sezione
Romana 9, no. 2 (1967): 177.
CHAPTER 15
SINGULARITY (1962–1966)
In the late spring of 1962, Stephen Hawking was facing the most important
academic examination of his life. Most undergraduates at Oxford didn’t have
to go through the ordeal of the viva, an oral exam akin to a thesis defense—it
was reserved for special or difficult cases. Stephen Hawking happened to be
a special and difficult case.
The three undergraduate years at Oxford were structured around the
“tutorial.” Each student would be assigned to a tutor—a faculty member in a
relevant field who would meet with a small group of undergraduates once or
twice a week. The undergrads would have to go to lectures and complete
weekly assignments, but those assignments were for the student’s own
benefit. All that really mattered, gradewise, were the final exams.
Hawking’s tutor, Robert Berman, was a physicist—a thermodynamicist
who studied how heat flowed through exotic solids. Berman recognized that
Hawking was brilliant, “completely different from his contemporaries.” But
for all his brilliance, Hawking had been an indifferent student—even lazy.
He thrived in theoretical physics, but he found it easy; he had to expend little
effort to get through the problem sets even as his fellow undergrads
struggled. “He could do any problem put before him without even trying,”
Berman told a reporter in the 1990s. But Hawking had little patience for
anything else; experimental physics, much less anything having to do with the
history of science, held no interest.1
During one summer, Hawking worked at the Royal Greenwich
Observatory helping the Astronomer Royal make measurements on double
stars and found the whole exercise dreadfully disappointing. Nothing at
Oxford excited any passion in him, and as a consequence, Hawking spent as
little effort on his studies as he possibly could. “At that time, the physics
course at Oxford was arranged in a way that made it particularly easy to
avoid work,” Hawking later wrote, and he avoided work with vigor,
spending a mere one thousand hours or so (by his own estimate) on his
studies during his entire time as an undergraduate. Indeed, working hard
would have been a source of shame rather than of pride, the mark of a “gray
man” rather than a lively fellow. Underneath the bravado, though, there was a
creeping ennui, “an attitude of complete boredom and feeling that nothing
was worth making an effort for.”2
Because Hawking didn’t feel challenged by the work his tutor gave him,
he didn’t exercise his mathematical muscles very well; he didn’t know it yet,
but he was actually a bit weak in the mathematical techniques he would need
to succeed in his chosen profession. For Hawking had decided to be a
cosmologist. But he had to get through the finals first, and he found it more
difficult than he had expected.
“Because of my lack of work, I had planned to get through the final exam
by doing problems in theoretical physics and avoiding questions that
required factual knowledge,” Hawking said in the Brief History of Time
documentary. “I didn’t do very well. I was on the borderline between a first-
and second-class degree, and I had to be interviewed to determine which I
should get.” This was the viva.3
Hawking’s future was resting on the outcome of the viva: he had been
accepted to Cambridge University, but that acceptance was contingent upon
getting a first-class degree. Without it, his career path was very much in
doubt—at one point, he even interviewed for a position with the Ministry of
Works, but then forgot to show up for the civil service examination.4
The professors at the viva knew how much was at stake for the young
Hawking. “They asked me about my future plans,” Hawking recalled. “I
replied, if they gave me a first, I would go to Cambridge. If I only got a
second, I would stay at Oxford. They gave me a first.”5
In his last year at Oxford, it was not at all clear that Hawking would ever
amount to anything. He was brilliant, to be sure—he had an extraordinary gift
for physics, and he could instantly see things that were obscure to others who
had expended much more effort. But expending effort didn’t much appeal to
him.
He would soon be a different man.
He would be diagnosed with an incurable disease and be given only a
few dozen months to live. He would find love. He would find a purpose. And
he would take his first few steps down a singular path.
Shortly after birth, a newborn baby begins trying to exert control over its
little body. The struggle goes from top to bottom. At first, she can’t even hold
her head upright; when her father scoops her up, he must take care to cradle
the head so that it doesn’t hang limply. Within a few weeks, the infant is able
to hold her head up, and then turn over in bed. A month or two later, the baby
can sit up, and, once she has control over her larynx, begins to babble rather
than merely cry. Soon, she learns to crawl, and the new parents realize how
much more effort it takes to care for a mobile child than an immobile one. A
few more weeks, and the baby is pulling herself upright. All too soon, she is
able to balance herself on her feet and begins to take a few unsteady steps,
trying to keep her oversized head and body upright supported only by a pair
of absurdly stumpy legs. And then, a year and change after the process has
begun, the child—no longer an infant, but a toddler—has mastered the ability
to balance, to walk, to run. From head to toe, she has tamed her muscles and
taught them to obey.
To many patients with amyotrophic lateral sclerosis—or motor neurone
disease, as it’s called in the United Kingdom—the disease is an unraveling, a
gradual loss of obedience of the muscles tamed in the first year of life. For
reasons yet poorly understood, the long nerve cells that transfer
electrochemical messages from the brain to the spine and the spine to the
muscles wither and die. The legs, the arms, the fingers, the larynx, all these
are cut off from the brain’s command, and, by default, disobey the orders that
they never receive. The denervated muscles slowly atrophy as they wait in
vain to be put to use, and the patient gradually wastes away.
By the end of his life, ALS had claimed almost all of Hawking’s muscular
control; he could not even move his fingers enough to click a button, but
instead had to control his computer by twitching his cheek. His head lolled
uncomfortably to the side unless propped up by a nurse, and a rivulet of
saliva typically rolled down his chin. Play the progress of his disease
backward through time and it’s remarkably similar to the toddler gaining
control, painstakingly re-created over the course of half a century. He gains
control over his limbs and is able to operate his computer and his
wheelchair; he regains his voice and speaks more and more intelligibly; his
fingers uncurl and he regains the ability to scrawl his name, to write, to walk
with crutches. And in 1962, at the beginning of the process, the only sign of
the disease was a bit of clumsiness, an unexpected fall here and there, and a
bit of trouble tying shoelaces.
In his last term at Oxford, Hawking was coming down a spiral staircase
when he slipped. “He kind of bounced all the way to the bottom. I don’t
know whether he lost consciousness, but he lost his memory,” said fellow
Oxford undergraduate Gordon Berry a number of years later. At first he was
a total amnesiac—unable to remember who he was. Over the course of a
couple of hours, his memory came back. Worried that he had damaged his
intellect, Hawking decided to take the test for Mensa, the “genius” society.
There was apparently no lasting damage. “He came back delighted that he
was able to get into Mensa. Absolutely delighted.”7
Stephen didn’t tell his mother, Isobel, about his increasing clumsiness: “I
think he began to notice that his hands were less useful than they had been,
but he didn’t tell us,” she later said. But he couldn’t keep the secret for long.
During the Christmas break home from his first term at Cambridge, Stephen
went skating with his family. Stephen was skating about, and then Isobel saw
him tumble helplessly to the ice. “He fell and he couldn’t get up. So I took
him to a café to warm up, and he told me all about it.” Isobel took Stephen to
see the family doctor, who referred Stephen to a specialist. In the meantime,
it was life as usual.8
New Year’s Day, 1963. Stephen’s friend Basil King threw a little party,
and Stephen was there in a black velvet jacket and bow tie, joking with his
friends. Basil’s younger sister, Diana, had invited her friend Jane Wilde.
Stephen and the eighteen-year-old Jane hit it off—she found his self-
deprecating humor delightful and decided she could overlook his long
fingernails and mop of floppy hair. And there was something about his eyes
—soulful gray eyes—and his broad smile that won her over. They exchanged
names and addresses, and Jane soon received an invitation to Stephen’s
twenty-first birthday party, on January 8.9
Stephen’s birthday was a fairly dreadful affair. Even though she knew
many of the people in attendance—Diana was there as well—Jane was
uncomfortable. The house was cold, the conversation stilted, and much of the
evening was spent with the guests challenging each other with brain-teaser
riddles. At that point, Jane didn’t see much of a future with Stephen—and
vice versa.10
Soon after his birthday, Stephen was admitted to St. Bartholomew’s
Hospital in London—where his father, Frank, had gone to medical school—
to undergo a battery of neurological tests. His frugal parents were willing to
spend for a private room, but Stephen refused; his “socialist principles”
wouldn’t allow him to use capital for that purpose. He was there for two
weeks, undergoing a series of increasingly unpleasant examinations to
determine what was wrong with him. Doctors took a muscle biopsy to see if
the problem might lie with the muscles themselves rather than the nerves.
They did an electromyographic study, sending shocks of electricity down his
neurons to test how well they conducted signals and transmitted them to the
muscles. And they did a procedure known as a myelogram to look at the
spine and spinal canal: as Hawking described it, doctors “injected some
radio-opaque fluid into my spine and with x-rays watched it go up and down
as they tilted the bed.”11
At the end of the ordeal, the doctors didn’t give Hawking a diagnosis,
telling him instead that he was an atypical case, and that his problems would
continue to get worse. Hawking wrote that he didn’t know that the culprit
was ALS until a few years later,iv but he knew he had been struck with an
untreatable, progressive ailment—a terminal disease. “The doctors offered
no cure and gave me two and a half years to live,” he later said. All they
could suggest is that he go back to Cambridge and try to finish his PhD.12
Hawking quite naturally fell into a depression. “[I] thought I would be dead
in a few years. There didn’t seem to be any point in carrying on,” he told one
early interviewer. Instead, he spent most of his time in his room listening to
Wagner and “drinking a fair amount”—an account Hawking tried to distance
himself from later on. When a reporter from Playboy asked him a number of
years later about his drinking binge, Hawking responded, “It’s a good story,
but it’s not true.… I took to listening to Wagner, but the reports that I drank
heavily are an exaggeration. The trouble is, once one article said it, others
copied it, because it made a good story. Anything that appeared in print so
many times has to be true.”13
On top of that, Hawking’s studies weren’t going well. Even before his
diagnosis, he was floundering and in danger of failing out. “At first, I was
doing very little work. I had very little mathematical background, so that
made it difficult to make any progress,” Hawking told a journalist two
decades later. His condition was getting worse day by day. “It seemed to be
developing very rapidly at first,” he said, “and I was very depressed. I didn’t
think there was any point in doing any research, because… I didn’t feel I
would live long enough to get my Ph.D.”14
Dennis Sciama watched his student with concern, but it would be up to
Stephen to pull himself out of his rut. Frank Hawking paid a visit to Sciama
to pressure him to let Stephen graduate early—presumably on humanitarian
grounds—but Sciama refused to bend the rules. If Stephen were to earn a
PhD, it wouldn’t be due to any special favors; he would have to earn it.15
In fact, despite Hawking’s disappointment at not being apprenticed to
Hoyle, Dennis Sciama was a boon for the young student. Like Hoyle, Sciama
had gotten his PhD from Paul Dirac, and he, too, was an advocate of steady-
state theory. But Sciama was less ambitious—and less fractious—than the
irascible Hoyle. He devoted most of his working hours to his students, and
like John Archibald Wheeler, he spent an enormous amount of time and effort
helping them launch their careers. Even if Sciama had been able to bend the
rules for a dying student, he probably sensed that Stephen needed to succeed
despite—not because of—his disease. Otherwise, any triumph he might
achieve would be hollow. But in early 1963, that hardly seemed a kindness.
The one thing that was going well for Stephen was his love life. In
February, he ran into Jane Wilde on a train platform. She had heard about his
diagnosis, but he refused to talk about the matter. So the two discussed more
pleasant things, and she accepted his invitation for a date. The two were soon
courting, which meant bopping back and forth variously to Cambridge or to
the outskirts of London on weekends. Usually by train. When they went by
car, Jane hung on for dear life; Stephen apparently drove like a man with
nothing to lose.v By the end of the spring semester, it seemed like they were a
couple.16
However, Stephen’s courtship, like his disease, progressed in fits and
starts. Jane went to Spain for the summer, and upon her return, Stephen didn’t
make much of an effort to track her down; he only contacted her in November
when he had come down to London for other reasons. And when Jane left for
a term abroad in Spain in the spring of 1964, Stephen didn’t bother to answer
any of her letters. Upon her return, she found him “terse and
uncommunicative.” On occasion, he exhibited such “hostility and frustration”
that Jane suspected he was “deliberately trying to deter me from further
involvement with him. It was too late. I was already so deeply involved with
him that there was no easy or obvious way out.” She was almost relieved
when the two decamped to different countries for the summer holiday.17
In the middle of 1964, Jane perceived a young man dogged by depression,
cynical and disheveled. However, in many respects, Stephen’s situation had
brightened considerably since the dark winter of his diagnosis two years
earlier. After a steep decline, the disease loosened its grip, and his condition
stabilized; he now had to walk with a cane, but it seemed quite certain that he
would defy the dire two-and-a-half-year sentence the doctors had passed
upon him. And after two years of flailing about, trying to master the
mathematics of the Einstein field equations, he finally got it—and he had a
distinct talent for the work. He had chosen his field wisely. It was around that
time that Dennis Sciama tried to get his young student’s brain into gear,
saying something to the effect of: “Well, you’re not dead yet. So, are you
ready to work on that problem I suggested?”18
He was. Hawking was primed to attack a thesis-worthy problem in
cosmology, but finding that problem was itself a major task. None of the
problems Sciama was suggesting were sticking. “It’s very difficult in these
advanced fields to find a good thesis topic. Cosmology at that time was a bit
fragmentary. It was not easy to say, ‘Well, here’s a problem. It will take you
three years; now go on and do it,’” Sciama said two decades later. “So
indeed [Hawking] looked around for quite a bit. While he did quite
interesting things, the real measure of his ability was hardly emerging yet,
and I can imagine it was a bit frustrating for him.” And Hawking, for his part,
didn’t think much of Sciama’s proposed topics for a thesis. “He was always
stimulating, but I didn’t agree much with his ideas,” Hawking later recalled.
But once out of his depression, Hawking’s mind was primed, just awaiting
the tinder to set it ablaze.19
The first sparks came in early 1964. Hawking shared an office at
Cambridge with a young postdoc—Jayant Narlikar—who had studied under
Hoyle and was busy hammering out some steady-state-related theories with
his former adviser. Specifically, Narlikar and Hoyle were trying to create a
new mathematical description of gravity in hopes that their version would
help them support the idea of an eternal universe. Hawking, quite naturally,
was interested in what his officemate was working on—and Narlikar was
more than willing to share. Narlikar even provided an early draft of a key
paper that he and Hoyle were about to submit to peer review. Hawking
played around with the equations on his own and found a flaw in their theory.
He realized that their idea wouldn’t work because certain key values
“diverged”—they went off to infinity, rendering the whole calculation
meaningless. But rather than telling Narlikar about the problem, Hawking
attended a meeting of the Royal Society in London where Hoyle was
presenting the idea to the scientific community for the first time.20
There were about a hundred scientists in attendance, and when Hoyle
finished his talk, he asked if there were any questions. Hawking rose slowly
to his feet and balanced himself with his cane. A hush fell over the crowd as
the unknown twenty-two-year-old graduate student had the temerity to
challenge the nation’s most famous astrophysicist:
Since this was the first time that Hoyle had presented his theory to the public,
everyone assumed that Hawking had done the calculations in his head, on the
spot during the lecture. Nobody suspected that he had seen a draft of the
paper beforehand, and had come to the meeting forearmed with what would
be the fatal blow to Hoyle-Narlikar gravitation. “Hoyle was furious,”
Hawking later wrote.22
Hoyle’s fury was irrelevant; Hawking was right. The values diverged.
The young upstart had gone head-to-head with the most famous astrophysicist
in Britain and come out on top. Not just that, he cultivated the impression that
he had done it off the cuff—that the eccentric-looking young gentleman
leaning on a cane harbored a mind so vast that it could do calculations in
minutes that Hoyle and his graduate student couldn’t figure out how to do in
months. It was a Feynman-quality performance. And, as it turned out, it was
Hawking’s first publishable-level research.
Knocking down someone else’s theory—especially a brand-new one that
hadn’t received much scrutiny—isn’t itself a huge feat, but it is real physics
nonetheless. Hoyle and Narlikar had submitted their calculations to the
Proceedings of the Royal Society A, where they were published at the
beginning of the summer; Hawking’s counterstrike, proving that the Hoyle-
Narlikar calculations were a dead end, was almost guaranteed a spot in the
same journal. But as with many other society journals, Hawking needed to
find a sponsor—in this case, a member of the Royal Society—who was
willing to transmit the paper to the editors of the journal. So he approached
Hoyle collaborator Hermann Bondi, whose lectures Hawking had attended a
few times. Bondi said yes; he probably felt honor-bound to help air any
substantive critiques of steady-state-related work—but the whole affair
probably didn’t leave Bondi with a great love for Hawking. That didn’t
matter… at the moment. For in October 1964, Hawking submitted his first
peer-reviewed paper to a prestigious journal; it would be published in the
summer of the following year.
The takedown of the Hoyle-Narlikar theory also became Chapter 1 of
Hawking’s PhD thesis. By itself, it wasn’t enough to merit a degree. But it
was a solid little piece of research, and Hawking was finally producing
something of merit. Despite the death sentence hanging over him, he had
transformed himself into a physicist.
Jane likes to think she broke Stephen out of his depression and gave him the
will to live. “We were going to defy the disease. We were going to defy the
doctors,” she told an interviewer in 2013. “And we were going to challenge
the future.” Stephen, too, cleaved to this narrative, giving Jane credit for his
renewed sense of purpose. “The disease wasn’t progressing so rapidly, and I
got engaged to be married. So that was really the turning point,” he told an
interviewer. “I realized that if I was going to get married, I would have to do
some work—I’d have to get a job. About that time, I began to understand
what I was doing as a mathematician.”23
However, in mid-1964, when Hawking was busy demolishing the Hoyle-
Narlikar theory, it wasn’t at all clear that the romance would survive. Their
parting at the end of the spring term left Jane wondering if they would see
each other when they both returned to Britain at the end of summer break.
And then, as Jane was traveling through Italy with her family, she received a
postcard from Stephen. “I was overjoyed to receive such an unexpected
piece of correspondence,” she later wrote.24
“Could Stephen really have been thinking of me as I had been thinking of
him?” she wondered. “It gave me grounds for daring to hope that he was
looking forward to seeing me at the end of the summer.” He was, in fact.
Stephen proposed to Jane in October 1964, at almost the same moment that
Hermann Bondi transmitted Hawking’s paper on Hoyle-Narlikar gravitation
to the Proceedings of the Royal Society.
Jane said yes.
She would wed a young man who wasn’t expected to live to the end of the
decade. As Jane put it in 2015: “[I] loved Stephen and wanted to do my best
for him. So I thought that I could easily devote two years of my life to help
somebody I loved—someone who had so much potential—achieve his
ambitions.” Neither she nor Stephen dreamed at the time that it would not be
two years of devotion, but twenty-five.25
Even early on, Jane got the sense that it wasn’t going to be easy to live
with Stephen. He had a tendency to assert dominance in a crowd by playing
the contrarian—picking fights with innocent bystanders. Jane herself was by
no means spared: “With me he would argue that artificial flowers were in
every way to be preferred to the real thing and that Brahms, my favourite
composer, was second-rate because he was such a poor orchestrator.” He
even embarrassed her in front of Alan Deyermond, her PhD adviser, with a
tirade about the uselessness of studying medieval literature. “You shouldn’t
take it personally,” he told her when she took offense. Yet she loved him, and
she was committed to helping him for the rest of his life, tragically
foreshortened though it might be.26
The upcoming nuptials forced Stephen to think more carefully about his
future, at least in the near to medium term. With a wife who wanted to
continue her studies, he was going to have to be the breadwinner, and that
meant finishing up his PhD and getting a fellowship of some sort. Luckily, his
research was finally going well; in late 1964 and early 1965, Sciama and his
group of acolytes—Stephen often among them—regularly took the train to
physics lectures in London. The Penrose talk in January, faithfully recounted
by Brandon Carter, provided that final spark, that one big idea, worthy of a
PhD thesis and more. The singularity theorem he had in mind was good
enough to fuel the start of his career. But he still had to find a job.
At the start of a chilly weekend in February 1965, Hawking eagerly
awaited Jane’s arrival. When she came through the door, left wrist in a cast,
he was horrified—horrified because he needed her to type up his fellowship
application. “I must admit that I was less sympathetic than I should have
been,” Hawking noted in his autobiography. Jane spent the entire weekend
writing it out longhand.27
It almost came to nothing. The application required two
recommendations, and Hawking had chosen to approach Hermann Bondi.
“After a lecture in Cambridge, I asked him about providing a reference, and
he looked at me, and in a vague way said yes, he would. Obviously, he didn’t
remember me, for when the college wrote to him for a reference, he replied
that he had not heard of me.” Sciama immediately got on the phone. Hawking
soon had a glowing recommendation.28
He got the fellowship. It would start that fall.
Jane and Stephen were married on July 15, a mere five days after the end
of the International Conference on General Relativity. At that conference,
Hawking presented his singularity theorem to some of the most important
relativists of the day, including John Archibald Wheeler and his students Kip
Thorne and Charles Misner. Hawking made quite an impression on them, and
he wound up fast friends with Thorne and Misner (who would be the
godfather to his son Robert two years later). And then it was off to an all-
too-brief stay in Suffolk before packing off to the Cornell summer school a
week later.
Though Jane was now married to Stephen, it was beginning to dawn on
her that Stephen was now married to physics.
Footnotes
i For a deeper explanation of Hubble’s work, see my Alpha and Omega,
chap. 3.
ii George Gamow and Isaac Asimov were the only people, at least in the
English language, who could even come close, at least until Carl Sagan
claimed the title in the late 1970s.
iii Steady-state theory was on the respectable end of the spectrum, but Hoyle
clung to it long after it had given up the ghost. On the wackier side of the
spectrum was the idea that oil had not come from dead organic material but
was being created continuously in the Earth’s core; another was that the
sunspot cycle was linked to flu pandemics, because influenza viruses came
from outer space.
iv If this is true, it was almost certainly willful ignorance; there’s little
chance that Frank would have been content until he had an affirmative
diagnosis in hand. As Stephen admitted in his autobiography, he didn’t press
the doctors for details at the time because he knew the details were bad.
v Even long after Hawking himself lost the ability to be behind the wheel,
this reckless-driving habit continued whenever he had any degree of control
over a car. Marika Taylor remembers driving her adviser around California
in a minivan in the late 1990s: “Stephen told me to do a U-turn, and I’m like,
‘But it says no U-turn!’ So he’s like, ‘No, you’re doing a U-turn.’ So I do a
U-turn.” Naturally, the rearview mirror filled with flashing red and blue
lights, and the police pulled the van over. Taylor adds, “They open up the
back of the van and say, ‘Oh, hello, Dr. Hawking! It’s you again.’”
CHAPTER 16
YLEM (1942–1962)
In front of a stone wall, eroded with the passage of time, some three dozen
members of the Oxford University boat club goof around as they pose for a
photograph. Four chaps sit on a sofa while a fifth stands on his head,
supported by grinning boys surrounding him. Several sit on or dangle from
the wall in various states of precariousness. One holds a hammer, seemingly
ready to strike the unsuspecting student in front of him; another proudly
displays a large letter “Y,” possibly the trophy of some sort of illicit raid of a
rival college. Some stare at the camera, some look away. Some are in tidy
suits, some are disheveled, one is entirely shirtless, and three do their best
Lawrence of Arabia imitation—keffiyahs and all. It’s a mishmash of chaotic
action.1
But one young man on the far right, near the edge of the frame, dominates
the tableau. He stands as rigid as a statue, left arm clutched to his chest and
right arm upraised with a kerchief in his hand—head thrown back, grimacing
in mock solemnity. That young man is Stephen Hawking.
Hawking seems younger than most of the others in the boat club; many of
the young men had been conscripted into the military before coming to
Oxford. Hawking was fortunate in that compulsory military service was
abolished in the United Kingdom in 1960, just as he would have been called
up.
The boat club offered some comradeship to the young Hawking, who was
not only bored at Oxford, but lonely. Not being the brawny type, he was
drafted to be a coxswain. Hawking would sit in the front of the boat in a
white straw hat, telling the eight rowers in the crew when to stroke and
where to steer. He was good at the job, but Norman Dix, the Oxford boatman
—the official in charge of the crews—was torn about whether or not to put
Hawking in charge of the best crew on the river. “The question always with
Stephen was should we make him coxswain of the first eight or the second
eight,” Dix later said. “Coxes can be adventurous. Some coxes can be very
steady people, you see. He was an adventurous type. You never knew quite
what he was going to do when he went out with a crew.”2
Oxford University, night. The river Thames flows silently in from the
northwest and, joining the Cherwell just south of campus, streams away to
the south and east. All’s quiet. On one of the footbridges spanning the river,
Hawking and a friend try not to snicker as they use a rope to suspend a plank
from the bridge. They climb down with paintbrushes:
A few minutes later, just visible in the dark, were the words VOTE
LIBERAL in foot-high letters along the side of the bridge and clear to
anyone on the river when daylight broke.
Then disaster struck. Just as Hawking was finishing off the last
letter, the beam of a flashlight shone down on them from the bridge and
an angry voice called out, “And what do you think you’re up to then?”
It was a local policeman.3
Hawking’s friend jumped off the plank, made it to the riverbank, and scurried
away. But Hawking was nicked. He was hauled off to the local constabulary
and given a good talking-to, but nothing more came of the incident.
Given the potent mix of boredom and beer—spiced with more than a
touch of arrogance—it’s amazing that Hawking didn’t get into more trouble
than that. “At Oxford, where he spent a lot of time drinking with oarsmen,”
wrote an old friend of his, “he seemed angry and frustrated in an unfocused
way, and his manner was provokingly raffish.” He was lucky enough to have
a tutor, thermodynamicist Bobby Berman, who wasn’t offended by the young
man’s affect, even when his peers were aghast at Hawking’s attitude. “He
used to produce his work every week for tutorial, and, as he never kept any
notes or papers or that sort of thing, on leaving my room, he would normally
throw it in my wastepaper basket,” Berman told filmmaker Errol Morris.
“And when he was with other undergraduates at the tutorial, and they saw
this happen, they were absolutely horrified. They thought he did this work in
probably half an hour. If they could have done it in a year, they wouldn’t have
thrown it in the wastepaper basket, they would have put it in a frame up on
their walls.”4
But Berman couldn’t engage Hawking, who spent most of his time coxing
on the river, playing bridge or poker or darts, and drinking port or beer—and
as little time as possible on physics. Even so, he quickly outshined the three
other undergraduates at Oxford in his year. One of them, Derek Powney, told
of one particularly hard problem set—thirteen problems in electricity and
magnetism—all of which seemed almost totally intractable. “I discovered
very rapidly that I couldn’t do any of them,” Powney said. He teamed up with
one of the students, and by the end of the week, as the problem set was
coming due, the two of them together had completed one and a half problems.
The third student, working alone, managed to finish one. “Stephen as always
hadn’t even started.” Hawking then went to his room to do the assignment,
and emerged about three hours later. “‘Ah, Hawking,’ I said, ‘how many have
you managed to do, then?’” Powney recalled. “‘Well,’ he said, ‘I’ve only had
the time to do the first ten.’ I think at that point we realized that it’s not just
that we weren’t on the same street, we weren’t on the same planet.” He
added, “It’s quite difficult to live all the time with people who are a lot
stupider than you are. I think that therefore you tend to become a very private
person, and to build almost a caricature of yourself as a defense.”5
Frank Hawking ponders a book, his head on his hand. Despite the round-
rimmed glasses, tweedy jacket, and mop of wispy gray hair—the very picture
of an academic type—there’s something steely about Frank. Where his son is
lanky, Frank is broad and robust, clearly used to the outdoors. There’s more
severity than warmth in his visage.6
Another photo, from a few years earlier, shows him on a rocky shoreline
in a khaki shirt and toting some sort of field kit over his shoulder. He gives a
half smile to the camera as he cups something indistinguishable—some
specimen, perhaps—in his right hand. One would guess him to be an
archaeologist rather than a medical doctor.7
Frank’s alma mater was Oxford, and he desperately wanted Stephen to
attend as well. But Stephen’s performance at St. Albans School was far from
stellar—despite his genius, Hawking always wound up in the bottom half of
his class. (“It was a very bright class,” Hawking liked to joke.) This made
Oxford very unlikely for Stephen, even without the additional worry of
obtaining a scholarship. So Frank decided to take matters into his own
hands.8
In the early spring of 1959, just before Stephen was to take the entrance
examinations for Oxford, Frank and Stephen paid a visit to Bob Berman,
Stephen’s prospective tutor. Frank reportedly applied pressure, enough that
Berman would ordinarily have been ill disposed to the candidate. But
Stephen scored so high on the exams that his father’s meddling—and
Stephen’s mediocre high-school grades—were moot. Hawking would go to
Oxford to study physics.9
Hair slicked back and in a dapper slate-gray suit, Billy Graham addresses a
sellout crowd at Harringay Arena in London, kicking off a three-month
evangelical “crusade.” “It’s been a long time since evangelism and revival
and Christ and God was front-page news around the world,” he exclaimed.
“And we thank God for it!” In 1954, Graham’s visit, full of American-style
evangelism, backed with a chorus of two thousand singers, was guaranteed to
set British tongues a-wagging.23
It even landed a convert or two. One of the boys in Hawking’s circle of
friends caught the spirit from Graham, and soon the boys of St. Albans were
chatting about matters theological.24 Hawking watched his schoolmates’
sudden zeal with some bemusement. As Michael Church, a boyhood friend,
described it:
We began talking about life and philosophy and so on, which I thought
I was very hot on at the time, and so I held forth.
Suddenly I was aware that he was egging me on, leading me to
make a fool of myself. It was an unnerving moment. I felt looked down
upon from a great height. I felt that he was watching, amused and
distant.… And there was some overarching arrogance, if you like,
some overarching sense of what the world was about.25
It must have irked his peers when Stephen won his school’s divinity prize
one year. As his mother, Isobel, put it, “It’s not surprising because his father
used to read him Bible stories from a very early age and he knew them all
very well.”26
For a brief moment, Hawking got caught up in an ESP craze, boosted, in
part, by the work of J. B. Rhine, a parapsychologist at Duke University who
wrote a few popular books on the subject, including one that came out in
1953. Stephen scrutinized the Duke research and quickly came to the
conclusion that the work was bunk. “Whenever the experiments got results,
the experimental techniques were faulty. Whenever the experimental
techniques were sound, the results were no good,” he later told a journalist.
“People taking it seriously are at the stage I was when I was a teenager.”27
Science held more interest to Stephen than theology or parapsychology,
and his opinions were equally strong. When he heard that the universe was
expanding, he refused to believe it. “A static universe seemed much more
natural,” he said. “It could have existed and could continue to exist forever.”
He would be a PhD student before he realized that he was wrong and that the
universe was, indeed, expanding from a Big Bang.28
Even though he was obviously bright, Stephen refused to apply himself in
school. Isobel remembered confronting him about being third from the bottom
in his class: “So I said, ‘Well, Stephen, do you really have to be as far down
as that?’ and he said, ‘Well, not a lot of people [did] much better.’ He was
quite unconcerned.”29
Two young boys sitting on the edge of a dock; behind them, the clear water
revealing no color but that of the green hills in the distance. The smaller of
the two squints in the Mediterranean sun, all buck teeth and splayed ears and
grinning with boyish excitement. That’s the eight-year-old Stephen. Next to
him, a little larger, a little older, and a little cooler, William Graves scowls
at the camera, hiding behind a pair of sunglasses.35
It was the early spring of 1950, and Frank was away, traveling, as
always. One year it was India, another China, another Africa, another to the
swamps of the Americas. Stephen’s sister Mary remembers thinking that her
father—and, by extension, fathers in general—were “like migratory birds.
They were there for Christmas and then they vanished until the weather got
warm.” This time, when Frank disappeared for the season, Isobel decided to
take the rest of the family to visit her friend Beryl, on the island of
Majorca.36
Isobel and Beryl had met when the two were studying at Oxford—an
unusual thing for women to do in those days—and both were ardent leftists
(Isobel had joined the Young Communist League). Beryl had just married the
writer Robert Graves after a complicated love affair.37
While Stephen and the family had a “wonderful time,” he didn’t quite fit
into the Graves household. He sparred with his and William’s tutor—the poet
W. S. Merwin—who insisted that, each day, the boys were to read a chapter
of the Bible and write an essay about what they had read. Stephen realized
that it was busy work, as his tutor “was more interested in writing a play for
the Edinburgh Festival than in teaching us.” At the time, Graves was deep
into a mystical phase, busy deconstructing Greek and biblical myths in search
of the ur-stories hidden beneath. “So there was no one to appeal to,”
Hawking lamented.38
William was put off by the young visitor. “[Stephen’s] arrival rather upset
my routine,” William told an interviewer in 2018. “He was a year and a half
younger than me, which is an age for small boys. Also, my friends were the
village children and I was used to speaking to them in Mallorquin, which
Stephen didn’t speak—of course.” But the one who was most upset by
Stephen’s presence had to be Robert Graves himself.39
Robert had been badly injured in World War I, nearly succumbing to a
chest wound—one that he never truly recovered from. He suffered, too, from
what’s now called post-traumatic stress disorder. As he described it in his
memoir, Goodbye To All That:
Since 1916, the fear of gas had obsessed me: any unusual smell, even
a sudden strong smell of flowers in a garden, was enough to send me
trembling. And I couldn’t face the sound of heavy shelling now; the
noise of a car back-firing would send me flat on my face, or running
for cover.40
Stephen, maybe four years old, sits in front of a model train set—a sleek
steam engine with four cars—sitting on a small wooden track. He has a
toddler’s face, the cheeks of a baby, but there’s a depth to the expression in
his eyes that’s a bit startling in a boy that young.42
The train set was a gift from his father, the spoils of an overseas trip to
the Americas right after the end of the war. The family had lived in north
London and suffered the privations of rationing and random bombing from
German rockets, and now that the war was over, they began to settle into the
new normality of peacetime.
Frank and Isobel enrolled Stephen in the Byron House School. In his
memoir, Stephen recalled complaining to his parents that he wasn’t learning
anything. He had a point—he didn’t learn to read until the age of eight. “We
were more concerned with my husband’s brilliance rather than Stephen’s,”
Isobel later said. “Still, Stephen was a self-educator from the start, and if he
didn’t want to learn things, it’s probably because he didn’t need to.”43
The postwar years also saw the introduction of a spate of modern drugs
and compounds that could be used against tropical diseases. Frank
investigated how safe and effective some of these compounds were against
parasitic ailments like malaria, and he studied the life cycles of disease
parasites in laboratory animals. He had a spate of publications in the 1940s
—including in esteemed journals like Nature and the British Medical
Journal—and was soon promoted to head up the division of parasitology at
the National Institute for Medical Research. The family moved to St. Albans,
a little farther north, to reduce Frank’s commute. Unlike the London
neighborhood where Stephen had spent his first few years, St. Albans wasn’t
populated with academics and intellectuals. He, like the rest of the
Hawkings, would find it a bit more difficult to fit in.44
An infant Stephen, swaddled in a white knit blanket. His father holds him
and gazes at him intently through round-rimmed glasses. Stephen peers back
at him just as intently, studying his father with intense gray eyes.45
In the early 1930s, so the story goes, Albert Einstein was in Hollywood,
entertaining a visit by a friend, the comedian Charlie Chaplin. They were
enjoying some tarts baked by Elsa Einstein and idly chatting when Einstein’s
son turned to Chaplin. “You are popular,” he said, “because you are
understood by the masses. On the other hand, the professor’s popularity with
the masses is because he is not understood.”1
In 1919, Albert Einstein underwent a rapid metamorphosis, transforming
almost overnight into an international celebrity—from a mere theoretical
physicist into the archetype of scientific genius. From the start, that
transformation was founded largely upon unintelligibility. In introducing the
public to the theory of relativity, the New York Times took pains to convince
readers of its inaccessibility. Einstein’s book about relativity was touted as
“A Book for 12 Wise Men”: “When he offered his last important work to the
publishers he warned them there were not more than twelve persons in the
whole world who would understand it, but the publishers took the risk.” That
legend—and that number, twelve—was repeated over and over as Einstein’s
fame grew, even as the man himself denied it. “That the story of the twelve
men never died out, even though Einstein himself and others offered
disclaimers, shows the hold that this phrase had on the public mind,” one
historian writes. “It was probably the most important factor in the growing
fame of the theory of relativity.”2
Surrounded by his disciples, Einstein had become the bearer of the new
physics, superseding the old Newtonian testament. That the physics was
written in a mathematical language that few members of the public could
fully understand enhanced the awe surrounding Einstein rather than
diminishing it. People were less interested in the theory of relativity than
what it represented, just as they were less interested in the flesh-and-blood
Einstein than they were in him as a symbol of the triumph of the human mind.
It was a symbol so powerful that members of the public would fight to get
a glimpse of it. In 1930, a crowd of more than four thousand people trying to
see a film about Einstein rioted through the American Museum of Natural
History. The riot began not because a group of New Yorkers were desperate
to learn more physics; it was because they wanted to share in the vision of
someone destined to become immortal, someone whose fame would outlast
even that of the great Charlie Chaplin.
A scientist with that much hold over the public imagination is born
perhaps once a century, or even more rarely. Since the death of Einstein,
there is only one scientist who has come close to achieving that level of
fame: Stephen Hawking.
It’s extremely rare for a scientist to become a celebrity, much less achieve
public immortality. At any given moment, perhaps a dozen or two in the
world might have enough fans to be recognized on the street once in a while.
Those who do almost never gain their fans because of the science they’ve
done; more often than not, a scientist’s celebrity has little to do with the
quality of his or her research.
Each year, the names of the new Nobel laureates are forgotten within days
—if those names ever register at all. Even the extraordinarily rare double
Nobelists, like John Bardeen, don’t necessarily get a share of fame. The list
of former Lucasian chairs alone is filled with names that are instantly
recognizable to mathematicians and physicists—Paul Dirac of the Dirac
equation, Joseph Larmor of the Larmor frequency, Gabriel Stokes of Stokes’
theorem—but who never received much attention from the public while they
were alive, and are now almost totally unheard of outside of the scientific
community. Even scientists who had an absolutely pivotal insight don’t
automatically achieve fame. James Clerk Maxwell, Erwin Schrödinger,
Murray Gell-Mann, Ludwig Boltzmann, Niels Bohr, Ernest Rutherford… the
names of these physicists might trigger a glimmer of recognition in the sort
interested in science, but even when they were alive, few people on the street
would have known who they were or what they had accomplished. No, it’s
not scientific accomplishment that turns a physicist into a celebrity, but
something else.
The most common form of scientific celebrity comes from popularization.
In some sense, bringing the ideas of science to the public is an act of
translation, a clever use of metaphor and other literary tricks to make work
written in the language of mathematics accessible in the vernacular. A
popularizer is an intermediary between science and the public, a priest who
helps the laity commune with a higher knowledge. Generally speaking,
scientists are not expected to take on the mantle of this priesthood.i
Even so, there is a steady flow of scientists who attempt to popularize
scientific advances, including, occasionally, ones at the very pinnacle of their
field. Of these, a few each generation achieve some degree of celebrity. But
it’s usually not the sort of celebrity that lasts. Fred Hoyle was a household
name in the 1950s, but few today not interested in cosmology can recall his
name, let alone recognize his face. In the 1970s and 1980s, Carl Sagan was
arguably the most famous living scientist, yet he’s probably better known
today for his novel, Contact, than he is for anything he did having to do with
science or science popularization. Even high priests come and go, and soon
fade into obscurity.
The symbol of Stephen Hawking in the second half of the twentieth century
was just as much a product of the media as Albert Einstein’s in the first half.
In Hawking, the press had seemingly found someone who could match
Einstein—overmatch him, even—on all fronts.
Hawking’s science extended directly from Einstein’s, making him the
natural successor to the old prophet. This armed Hawking, like his
predecessor, to be one of the few who could gaze directly at the most
destructive forces in the universe; he descended, alone, into the maw of a
black hole and came back with esoteric knowledge that nobody else could
have retrieved. And he was purportedly on the path to complete Einstein’s
unfinished quest for a theory of everything: even though this never formed a
serious thread in Hawking’s scientific research (and he lagged far beyond the
string theorists and others who were working on unification), this quest was
a central part of the Hawking legend.
Hawking, like Einstein, was a man of the people—someone to be adored
rather than feared. Hawking’s sense of humor was wicked, and every bit as
self-effacing as Einstein. Einstein had tried to popularize his science.
Hawking had outdone him, writing a best-seller about his physics. And
whereas the public was disarmed by Einstein’s charming eccentricities, it
was transfixed by Hawking’s infirmity, finding it heroic as well as tragic. In
his shriveled body, he was a pure mind: a Tiresias whose crushing physical
disability was more than repaid by a divine gift of insight.
Little wonder that, like Einstein, Hawking was anointed prophet. But in
becoming a prophet, he would become a symbol. Unlike humans, symbols
can achieve immortality.
It took charm and discipline to maintain the simplicity of that image, even
as the complexity of Hawking the human being was visible just beneath the
surface. Hawking the symbol was arguably the most celebrated popularizer
of science in our time. Yet he was a man who had such difficulty
communicating that he had little choice but to save time and effort by
populating his speech and his writing with recycled phrases—when, in fact,
he himself was writing the prose that he set his name to. Hawking the symbol
was a secular saint, kindly and humble and, above all, harmless. Yet in
person, he could be stubborn, peevish, and proud. Divorced twice and with
periods of alienation from his children, even his family knew that he could be
a hard person to be around.
Hawking the symbol was a stoic who transcends life; his perseverance
was inspiring, even to—especially to—the people who knew him best. “I
was with him in the hospital a few times,” Kip Thorne recalls. “Once, he
was in a fairly bad shape, and the only way for him to communicate was by
using cards his carers would put up, and he basically indicated yes or no for
whether some letter on a card or some symbol on the card was what he was
trying to focus on,” Thorne says. “Even then, he didn’t show particularly
great frustration. It was quite incredible, really.” He never showed any self-
pity, and seldom even any vulnerability, to his family and friends, much less
to the public. At the same time, Hawking the person was dependent upon the
people around him to sacrifice themselves for his survival. And for his
hedonism. Hawking had a strong appetite not just for food and wine, but for
women—for extracting the pleasures he could from a life he thought could be
snatched away at any moment.7
As a scientist, Hawking the symbol was supposed to be a lone genius who
towered over his peers from his wheelchair, working to fulfill Einstein’s
dream of a theory of everything. The authentic Hawking was just one bright
star in a constellation of scientists in the United Kingdom, the United States,
and the Soviet Union who were transforming general relativity and
cosmology in the 1960s and 1970s; some of Hawking’s most important
contributions to physics were collaborative rather than solo achievements.
Hawking’s cardinal scientific achievement—the discovery of black-hole
radiation—was a triumph that used both quantum theory and relativity in a
physical regime where the two must give way to an overarching theory of
everything. Yet his work brought the world no closer to that theory, and by
the time he became an international celebrity in the late 1980s, his most
important scientific contributions were well behind him.
Hawking the scientist, who didn’t find the theory of everything, died in
2018; but Hawking the symbol, who brought us The Theory of Everything,
lives on. Just as Hawking worked to maintain his image during his lifetime,
the nonprofit set up by his family and friends—the Hawking Foundation—is
continuing to do so after his death.
More than that of any other human on the face of the planet, Stephen
Hawking had a life that was preserved in a digital medium. Every utterance
for the last few decades existed only through the mediation of a computer—a
computer that could, and did, store his conversations. Sitting quiescent in
those digital memory banks are, no doubt, traces of all the hidden
collaborations, the worries about money, the petty jealousies, his
uncertainties about his scientific legacy: all the rough edges of his humanity.
But the foundation has not made even Hawking’s paper archives available to
those who wish to see the human behind the symbol—despite a supposed
plan to donate at least some of them to public collections in order to avoid
the inheritance tax. Even a small cache of papers that were housed at a
library at Cambridge University were reclaimed and removed from display.
Only a biographer selected by the foundation is granted even limited access
to consult not just Stephen’s personal archives, but to get access to Jane,
Lucy, and the rest of the Hawking family.8
But it is in the messy humanity of the person rather than the pristine
symbolism of the image where the real tragedy and triumph of Stephen
Hawking are to be found.
The tragedy of Hawking’s life was not being diagnosed with a terminal
disease; nor was his triumph in overcoming that disease to become the
greatest physicist since Einstein. That is the story of the symbol. The human’s
story is much more complex.
To be sure, Hawking’s persistence and good humor in the face of his
deteriorating condition was inspiring, but to Hawking, that was hardly any
sort of triumph; it was merely survival. Nor was it false modesty when
Hawking rubbished the idea of his being a latter-day Newton, even though
there’s little doubt that he was a physicist of the first rank and made major
contributions to general relativity and cosmology. Hawking’s was not the
best mind since Newton and Einstein and Galileo, as he himself repeatedly
insisted.
Hawking’s triumph was to discover something new that changed our
understanding of the universe in important ways—and to do so multiple
times. More than that, Hawking’s triumph was also to act as a catalyst for
important ideas that were brewing in his field, and to inspire a generation of
scientists who wished to follow in his footsteps.
The central tragedy of Hawking the human being is also quite different
from that of Hawking the symbol. Hawking didn’t retreat into his mind as a
result of his disease. Since childhood, Hawking had been cerebral to the
extreme. Even when it wasn’t clear whether he would fail out of school, the
core of Hawking’s identity, of his self-worth, was the superiority of his
brain. It was what he always wanted to be known for.
Yet even as he achieved this goal, he was denied it, at least in his own
mind. Hawking always suspected that, to a large degree, his celebrity—and
even to some degree his academic success—rewarded his infirmity rather
than his brain.
One of the curious facts about Hawking was that he was unusually fond of
—even devoted to—a celebrity that, at first glance, seems to be his polar
opposite. Stephen Hawking absolutely adored Marilyn Monroe.
Errol Morris, the film director, was chatting with Hawking during the
filming of A Brief History of Time when the subject of Monroe came up. As
Morris put it:
Finally, I said, “I figured it out, why you have all these pictures of
Marilyn Monroe on the wall. Like you, she was a person appreciated
for her body and not necessarily her mind.”
And he gave me this really crazy look, like, “What the fuck are you
saying, Mr. Morris?” He gave me this crazy look, and then finally,
there’s a click, and he says, “YES.”9
Of all the paradoxes in Hawking’s life, this one may have been the most
profound.
Footnotes
i Sometimes, their reputations even suffer if they try. In 1992, Carl Sagan was
nominated to the National Academy of Sciences. He was rejected. “If he had
not done television, he probably would be in the academy,” one academy
member who was present at the vote later commented. More often, though, a
scientist’s attempt to commune with the public is met with mild scorn or mere
indifference from his or her peers.
ii Or at least its backside. See Exodus 33:21–23.
iii Even Einstein’s relativity—which was very much his brainchild—wasn’t
devised in a vacuum. Marcel Grossman, for example, was an important
collaborator who helped Einstein devise the mathematics necessary for the
general theory. In addition, mathematicians like Henri Poincaré and David
Hilbert were also having insights close to full-on discovery of the theory.
ACKNOWLEDGMENTS
Get sneak peeks, book recommendations, and news about your favorite
authors.
Virtual Unreality: Just Because the Internet Told You So, How Do You
Know It’s True?
Sun in a Bottle: The Strange History of Fusion and the Science of Wishful
Thinking
Alpha and Omega: The Search for the Beginning and End of the Universe
PROLOGUE
1. Jacqui Deevoy, “Has Stephen Hawking Been Replaced with a ‘Puppet’?,” Daily Mail, January 12,
2018, www.dailymail.co.uk/femail/article-5261939/Has-Stephen-Hawking-replaced-puppet.html.
2. Leonard Susskind, personal communication with author.
CHAPTER 2: RIPPLES
1. Maggie Fox, “Gravitational Wave Work Wins Physics Nobel Prize,” NBC News, October 3, 2017,
www.nbcnews.com/science/science-news/gravitational-wave-works-wins-physics-nobel-prize-n807081.
2. Sarah Lewin, “Nobel Prize for Physics: Einstein Would Be ‘Flabbergasted’ by Gravitational Wave
Win,” Space.com, October 3, 2017, www.space.com/38350-nobel-prize-physics-gravitational-waves-
einstein.html.
3. Kip Thorne, “Warping Spacetime,” in The Future of Theoretical Physics and Cosmology:
Celebrating Stephen Hawking’s 60th Birthday, ed. G. W. Gibbons, S. J. Rankin, and E. P. S. Shellard
(Cambridge: Cambridge University Press, 2003), 4–5, reprinted by Kip S. Thorne, California Institute of
Technology, at www.cco.caltech.edu/~kip/scripts/PubScans/VI-47.pdf.
4. Thorne, “Warping Spacetime,” 30.
5. Stephen Hawking, The Universe in a Nutshell (New York: Bantam Books, 2001), 14.
6. See video accompanying Ian Sample, “Stephen Hawking Unveils Formulae for England World Cup
Success,” The Guardian, May 28, 2014, www.theguardian.com/science/2014/may/28/stephen-
hawking-formulae-england-world-cup-success.
7. “Ridley Turtle Tipped for Oily Exit,” Paddy Power press release, May 25, 2010.
8. See Charles Seife, Proofiness: The Dark Arts of Mathematical Deception (New York: Viking,
2010), 65–66; Sample, “Stephen Hawking Unveils Formulae.”
9. Sample, “Stephen Hawking Unveils Formulae.”
10. Sample, “Stephen Hawking Unveils Formulae.”
11. Al Zuckerman, personal communication with author.
12. J. H. Taylor and J. M. Weisberg, “A New Test of General Relativity: Gravitational Radiation and the
Binary Pulsar PSR1913+16,” Astrophysical Journal 253 (1982): 908.
13. B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), “Observation of
Gravitational Waves from a Binary Black Hole Merger,” Physical Review Letters 116, 061102 (2016).
14. “The view from GR,” presented at KITP Rapid Response Workshop, August 23, 2013,
http://online.kitp.ucsb.edu/online/fuzzorfire_m13/hawking/rm/jwvideo.html.
15. Charles Q. Choi, “No Black Holes Exist, Says Stephen Hawking—at Least Not Like We Think,”
National Geographic, January 27, 2014, www.nationalgeographic.com/news/2014/1/140127-black-
hole-stephen-hawking-firewall-space-astronomy.
16. S. W. Hawking, “Information Preservation and Weather Forecasting for Black Holes,”
arXiv:1401.5761v1, January 22, 2014, 3.
17. Gareth Morgan, “Stephen Hawking Says There Is No Such Thing as Black Holes, Einstein Spinning
in His Grave,” Express, January 24, 2014, www.express.co.uk/entertainment/gaming/455880/Stephen-
Hawking-says-there-is-no-such-thing-as-black-holes-Einstein-spinning-in-his-grave; Mark Prigg,
“Stephen Hawking Stuns Physicists by Declaring ‘There Are No Black Holes’—But Says There Are
GREY Ones,” Daily Mail, January 24, 2014, www.dailymail.co.uk/sciencetech/article-
2545552/Stephen-Hawking-admits-no-black-holes-GREY-holes.html.
18. Ramin Setoodeh, “Toronto: ‘The Theory of Everything’ Made Hawking Cry,” Variety, September 7,
2014.
19. Ramin Setoodeh, “How Eddie Redmayne Became Stephen Hawking in ‘The Theory of
Everything,’” Variety, October 28, 2014.
20. The Theory of Everything, directed by James Marsh, November 2014.
21. Catherine Shoard, “Stephen Hawking’s First Wife Intensifies Attack on The Theory of Everything,”
The Guardian, October 3, 2018.
22. Ben Travis and Alex Godfrey, “Eddie Redmayne on Meeting Stephen Hawking for The Theory of
Everything,” Empire, March 14, 2018.
23. For Jane Hawking’s film rights, see Film and Music Entertainment, Inc., CIK#0001309152, Form
10SB12G/A, 2006-02-06.
24. “The Theory of Everything,” Box Office Mojo, www.boxofficemojo.com/movies/?
id=theoryofeverything.htm.
25. Stem Cell Universe with Stephen Hawking, Discovery Science, February 3, 2014.
26. Errol Morris, personal communication with author.
27. Author’s analysis.
28. Morris, personal communication with author.
29. “Stephen Hawking recorded iNNOCENCE + eXPERIENCE monologue,” YouTube, posted May
17, 2015, by Bob Mackin, www.youtube.com/watch?v=YZM7uM8OyxA.
30. “Report and Financial Statements for the Period 1 April 2016 to 31 March 2017,” Stephen Hawking
Foundation, Charity number 1163521.
31. Zuckerman, personal communication with author.
32. BBC Radio 4 Today, March 18, 2014, www.bbc.com/news/av/science-environment-
26625791/stephen-hawking-wins-inflation-debate.
33. Dennis Overbye, “Space Ripples Reveal Big Bang’s Smoking Gun,” New York Times, March 18,
2014, A1.
34. Chris Havergal, “Cambridge University’s Stephen Hawking Claims Victory in Bet with Neil Turok
After Cosmic Wave Discovery,” Cambridge Evening News, March 18, 2014.
35. George P. Mitchell, “George Mitchell Lays Groundwork for New Texas A&M Science Initiative
with $35 Million Gift,” press release, Texas A&M University, November 3, 2005; “Hawking Honored
with Auditorium at Texas A&M,” Associated Press, April 6, 2010.
36. Andy Strominger, personal communication with author; S. W. Hawking, “The Information Paradox
for Black Holes,” arXiv:1509.01147v1, September 3, 2015.
37. Strominger, personal communication with author.
38. Strominger, personal communication with author. The elision is “information paradox with
supertranslation hair.” Supertranslation hair is one variety of soft hair, which is the subject of the first
of three papers by Hawking, Strominger, and Perry; for technical reasons, this variety of soft hair
cannot, in fact, solve the paradox.
39. Raphael Bousso and Massimo Porrati, “Soft Hair as a Soft Wig,” arXiv:1706.00436v2, September
20, 2017; Marika Taylor personal communication with author.
40. Lawrence M. Krauss (@LKrauss1), Twitter, September 25, 2015, 4:39 p.m.,
https://twitter.com/lkrauss1/status/647510799678750720.
41. Lawrence M. Krauss (@LKrauss1), Twitter, January 11, 2016, 10:46 a.m.,
https://twitter.com/lkrauss1/status/686574829542092800; “Thursday: Scientists to Provide Update on
Search for Gravitational Waves,” LIGO Media Advisory, February 8, 2016.
42. Dr Erin Ryan, (@erinleeryan), Twitter, February 11, 2016, 10:14 a.m.,
https://twitter.com/erinleeryan/status/697800782175997952.
43. “LIGO Detects Gravitational Waves—Announcement at Press Conference (part 2), February 11,
2016, www.ligo.caltech.edu/WA/video/ligo20160211v12.
44. Stephen Hawking, A Brief History of Time (New York: Bantam Books, 1998), 93–94.
45. LIGO Scientific Collaboration, “GWTC-1: A Gravitational-Wave Transient Catalog of Compact
Binary Mergers Observed by LIGO and Virgo During the First and Second Observing Runs,”
arXiv:1811.12907v2, December 16, 2018.
46. Kip Thorne, personal communication with author.
CHAPTER 3: MODELS
1. “Opening Ceremony—London Paralympic Games,” YouTube, posted August 29, 2012, by Paralympic
Games, www.youtube.com/watch?v=Kd4FgGSY5BY.
2. “Opening Ceremony—London Paralympic Games”; Stephen Hawking, My Brief History (London:
Bantam Books, 2018), 122.
3. Andy Strominger, personal communication with author.
4. John Boslough, Stephen Hawking’s Universe (London: HarperCollins, 1995), 93–94.
5. Stephen Hawking, A Brief History of Time (New York: Bantam Books, 1998), 13.
6. Edwin M. McMillan, “Current Problems in Particle Physics,” Science 152, no. 3726 (May 27, 1966):
1212, https://doi.org/10.1126/science.152.3726.1210; Willis Lamb, “Fine Structure of the Hydrogen
Atom,” Nobel Prize Lecture, December 12, 1955, www.nobelprize.org/uploads/2018/06/lamb-
lecture.pdf (quotation marks omitted).
7. “Stephen Hawking on Higgs: ‘Discovery Has Lost Me $100,’” BBC News, July 4, 2012,
www.bbc.com/news/av/science-environment-18708626/stephen-hawking-on-higgs-discovery-has-lost-
me-100.
8. Peter Guzzardi, personal communication with author; Hawking, My Brief History, 122.
9. Amber Goodhand, “Renowned Physicist Stephen Hawking Frequents Sex Clubs,” Radar Online,
February 24, 2012, https://radaronline.com/exclusives/2012/02/stephen-hawking-sex-clubs-physicist-
freedom-acres.
10. “A Brief History of the Time Stephen Hawking Went to a Sex Club: University Says Physicist
Visited California Swingers’ Club with Friends,” Daily Mail, February 27, 2012,
www.dailymail.co.uk/news/article-2106025/Stephen-Hawking-visits-California-swingers-sex-club.html.
11. “Freedom Acres Resort,” Freedom Acres Resort, http://freedomacresresorts.com, accessed May
30, 2019; Russ Thomas, personal communication with author.
12. Errol Morris, personal communication with author.
13. Jane Hawking, Music to Move the Stars: A Life with Stephen Hawking (London: Pan, 2000),
328.
14. “NMC Announces Fitness to Practise Hearing Outcome Involving Former Nurse of Professor
Stephen Hawking,” Nursing and Midwifery Council, press release, March 12, 2019.
15. Andrew Holgate, “The Trouble with Being a Genius: Stephen Hawking Has Had to Overcome
Extraordinary Obstacles in His Life. His Memoir Hints at His Travails [Eire Region],” Sunday Times
(London), September 22, 2013.
16. Chuck Leddy, “‘My Brief History’ by Stephen Hawking,” Boston Globe, September 10, 2013.
17. Robert P. Crease, “A Cosmological Life,” Nature 501, no. 7466 (September 12, 2013): 162.
18. Stephen Hawking, Black Holes and Baby Universes and Other Essays (New York: Bantam
Books, 1993), 1–39.
19. Martin Rees, “Stephen Hawking—An Appreciation,” March 14, 2018,
www.ctc.cam.ac.uk/news/Hawking_an_appreciation-Rees.pdf.
20. Stephen Hawking, “This Is the Most Dangerous Time for Our Planet,” The Guardian, December 1,
2016, www.theguardian.com/commentisfree/2016/dec/01/stephen-hawking-dangerous-time-planet-
inequality; Robert Booth, “Stephen Hawking: I Fear I May Not Be Welcome in Donald Trump’s US,”
The Guardian, March 20, 2017, www.theguardian.com/science/2017/mar/20/stephen-hawking-trump-
good-morning-britain-interview.
21. Stephen Hawking, “The NHS Saved Me. As a Scientist, I Must Help to Save It,” The Guardian,
August 18, 2017, www.theguardian.com/commentisfree/2017/aug/18/nhs-scientist-stephen-hawking.
22. Harriet Sherwood and Matthew Kalman, “Stephen Hawking Joins Academic Boycott of Israel,”
The Guardian, May 7, 2013, www.theguardian.com/world/2013/may/08/stephen-hawking-israel-
academic-boycott; Robert Booth and Harriet Sherwood, “Noam Chomsky Helped Lobby Stephen
Hawking to Stage Israel Boycott,” The Guardian, May 10, 2013,
www.theguardian.com/world/2013/may/10/noam-chomsky-stephen-hawking-israel-boycott.
23. Iain Thomson, “Israeli Activists Tell Hawking to Yank His Intel Chips over Palestine,” The Register,
May 9, 2013, www.theregister.co.uk/2013/05/09/israel_boycott_stephen_hawking_intel; Averil
Parkinson and Jonathan Rosenhead, “Professor Stephen Hawking and the Academic Boycott of Israel,”
BRICUP Newsletter 120, April 2018, 3.
24. Steven Plaut, “Hawking,” Zionist Conspiracy, May 8, 2013,
http://zioncon.blogspot.com/2013/05/hawking.html.
25. S. W. Hawking, “Virtual Black Holes,” Physical Review D 53, no. 6 (March 15, 1996): 3099–3107,
https://doi.org/10.1103/PhysRevD.53.3099; John Gribbin, “Hawking Throws Higgs into Black Holes,”
New Scientist, December 2, 1995, www.newscientist.com/article/mg14820062-400-hawking-throws-
higgs-into-black-holes.
26. Alastair Dalton, “Clash of the Atom-Smashing Academics,” The Scotsman, September 2, 2002.
27. Dalton, “Clash of the Atom-Smashing Academics.”
28. “On the Hunt for the Higgs Boson,” BBC, September 9, 2008,
http://news.bbc.co.uk/today/hi/today/newsid_7598000/7598686.stm.
29. Mike Wade, “Higgs Launches Stinging Attack Against Nobel Rival,” The Times (London),
September 11, 2008.
CHAPTER 5: CONCESSIONS
1. Judy Bachrach, “A Beautiful Mind, an Ugly Possibility,” Vanity Fair, June 2004,
www.vanityfair.com/news/2004/06/hawking200406. See also “Is Stephen Hawking Being Abused by
His Wife?,” People, updated February 9, 2004, https://people.com/archive/is-stephen-hawking-being-
abused-by-his-wife-vol-61-no-5.
2. Hawking acquaintance, personal communication with author.
3. Various, personal communications with author.
4. Bachrach, “A Beautiful Mind, an Ugly Possibility”; Olga Craig, “Is This the Brief History of a
Troubled Marriage?,” The Telegraph, January 25, 2004,
www.telegraph.co.uk/news/uknews/1452515/Is-this-the-brief-history-of-a-troubled-marriage.html.
5. Assistant, personal communication with author.
6. Natalie Clarke, “Professor Hawking in Assault Probe,” Daily Mail, January 2004,
www.dailymail.co.uk/news/article-206323/Professor-Hawking-assault-probe.html; “Is Stephen Hawking
Being Abused by His Wife?”; Craig, “Is This the Brief History”; Laura Peek, “Hawking Denies Assault
by Wife After Fresh Allegations by Carer,” The Times (London), January 24, 2004.
7. Leonard Mlodinow, Stephen Hawking: A Memoir of Friendship and Physics (New York:
Pantheon Books, 2020), 163.
8. Craig, “Is This the Brief History.”
9. David Sapsted, “Hawking Defends His Wife After Assault Claims,” The Telegraph, January 24,
2004, www.telegraph.co.uk/news/uknews/1452453/Hawking-defends-his-wife-after-assault-
claims.html.
10. Cambridgeshire Constabulary, “Professor Stephen Hawking,” press release, March 29, 2004.
11. Nurse, personal communication with author.
12. “Shadowland,” The Age, April 21, 2004, www.theage.com.au/entertainment/books/shadowland-
20040421-gdxpqh.html.
13. Kitty Ferguson and S. W. Hawking, Stephen Hawking: An Unfettered Mind (New York: St.
Martin’s Press, 2017), 286; “Stephen Hawking’s Grand Design: The Meaning of Life Full Episode,”
YouTube, posted September 16, 2015, by S. M. Taif- Ul-Kabir, www.youtube.com/watch?
v=usdqCexPQww.
14. “Shadowland.”
15. Max Planck, The Origin and Development of the Quantum Theory: Being the Nobel Prize
Address Delivered Before the Swedish Academy of Sciences at Stockholm, 2 June 1920, trans. H.
T. Clarke and L. Silberstein (Oxford: Clarendon Press, 1922).
16. Author’s personal recollection of July 21, 2004; Jenny Hogan, “Hawking Cracks the Paradox,” New
Scientist 183, no. 2456 (July 17, 2004), 11.
17. “The Hawking Paradox,” Horizon, Season 42, Episode 2, directed by William Hicklin, BBC, aired
September 15, 2005.
18. Leonard Susskind, personal communication with author, as described in Charles Seife, “A General
Surrenders the Field, but Black Hole Battle Rages On,” Science 305, no. 5686 (August 13, 2004): 934,
https://doi.org/10.1126/science.305.5686.934.
19. Wager posted at John Preskill’s Caltech website, www.theory.caltech.edu/~preskill/info_bet.html.
20. Charles Seife, “Hawking Slays His Own Paradox, but Colleagues Are Wary,” Science 305, no.
5684 (July 30, 2004): 586, https://doi.org/10.1126/science.305.5684.586.
21. Andy Strominger, personal communication with author.
22. S. W. Hawking, “Information Loss in Black Holes,” Physical Review D 72, no. 8 (October 18,
2005): 084013, https://doi.org/10.1103/PhysRevD.72.084013.
23. Scott Feschuk, “Briefer Madness,” Maclean’s 118, no. 46 (November 14, 2005): 148; Jim al-Khalili,
“Short Cut to Space-Time,” Nature 438, no. 7065 (2005): 159–159, 161.
24. Matthew Syed, “Only Partially Blinded by Science,” The Times (London), October 15, 2005.
25. Stephen Hawking, God Created the Integers: The Mathematical Breakthroughs That Changed
History (Philadelphia: Running Press, 2007), 291.
26. Stephen Gaukroger, Descartes: An Intellectual Biography (Oxford: Clarendon Press, 2002), 416.
27. Hawking, God Created the Integers, 675.
28. Des MacHale, George Boole: His Life and Work (Dublin: Boole Press, 1983), 155–156.
29. Gil King, personal communication with author.
30. “Jeff Koons,” Journal of Contemporary Art, October 1986, www.jca-online.com/koons.html.
31. Hawking, God Created the Integers, 1162.
32. John Moran, “‘Theory of Everything’ Not for Novices,” Chicago Tribune, June 13, 2002, 4–5.
33. Hillel Italie, “Hawking Tries to Stop Publication of Lectures,” Associated Press, in Journal Gazette,
May 24, 2002, 4W.
34. Scott Flicker, “Complaint Against New Millennium Entertainment,” April 26, 2002, addressed to J.
Howard Beales III, director of consumer protection, no. 2130979.
35. Michael Hiltzik, “It’s a Safe Bet That He’ll Sue in This Town Again,” Los Angeles Times, June 2,
2003, www.latimes.com/archives/la-xpm-2003-jun-02-fi-golden2-story.html.
36. “Latest News,” Stephen Hawking website, Internet Archive, Wayback Machine, June 9, 2003,
https://web.archive.org/web/20030609215606/http://hawking.org.uk/home/hindex.html.
37. Stephen Hawking, My Brief History (London: Bantam Books, 2018), 89; Court employee, personal
communication with author; Elaine Hawking, “Baptism Testimony of Elaine Hawking,” Internet
Archive, April 2018, https://archive.org/details/chipping-campden-baptist-church-513231/2018-04-01-
Baptism-Testimony-Elaine-Hawking-Elaine-Hawking-46453311.mp3.
38. Al Zuckerman, personal communication with author.
39. Hawking, My Brief History.
40. “Shadowland.”
41. “Stephen Hawking Calls for Mankind to Reach for Stars,” Space Daily, June 14, 2006,
www.spacedaily.com/reports/Stephen_Hawking_Calls_For_Mankind_To_Reach_For_Stars.html.
42. Lucy and Stephen Hawking, with Christophe Galfard, illustrated by Garry Parsons, George’s Secret
Key to the Universe (London: Doubleday, 2007), 4–5.
43. Edge, www.edge.org/3rd_culture/krauss06/images/35.jpg; Agency, “Stephen Hawking Pictured on
Jeffrey Epstein’s ‘Island of Sin,’” The Telegraph, January 12, 2015,
www.telegraph.co.uk/news/science/stephen-hawking/11340494/Stephen-Hawking-pictured-on-Jeffrey-
Epsteins-Island-of-Sin.html.
44. “Stephen Hawking Eulogizes George Mitchell as Science Visionary, World-Changer, Friend,” Texas
A&M College of Science (blog), August 9, 2013, https://science.tamu.edu/news/2013/08/stephen-
hawking-eulogizes-george-mitchell-as-science-visionary-world-changer-friend.
45. Jesse Drucker, “Kremlin Cash Behind Billionaire’s Twitter and Facebook Investments,” New York
Times, November 5, 2017, www.nytimes.com/2017/11/05/world/yuri-milner-facebook-twitter-
russia.html; Stephen Hawking, “Yuri Milner,” Time, April 21, 2016, https://time.com/collection-
post/4300003/yuri-milner-2016-time-100.
46. Milner met Hawking at a conference in Moscow in 1987. Yuri Milner, “Stephen Hawking: The
Universe Does Not Forget, and Neither Will We,” Scientific American (blog), March 29, 2018,
https://blogs.scientificamerican.com/observations/stephen-hawking-the-universe-does-not-forget-and-
neither-will-we; Marika Taylor, personal communication with author.
47. Marika Taylor, personal communication with author.
48. “Peter Diamandis: Stephen Hawking Hits Zero G,” YouTube, TED Talk, posted July 8, 2008, by
TED, www.youtube.com/watch?v=0VJqrlH9cdI; “Reimbursable Space Act Agreement Between
National Aeronautics and Space Administration, John F. Kennedy Space Center and Zero Gravity
Corporation for Recurring Use of the Shuttle Landing Facility,” signed March 20, 2006 (James
Kennedy), and March 30, 2006 (Peter Diamandis).
49. Diamandis TED Talk.
50. “Review of NASA’s Microgravity Flight Services,” NASA Office of Inspector General, Report no.
IG-10-015; “Move to New Planet, Says Hawking,” BBC, November 30, 2006,
http://news.bbc.co.uk/2/hi/6158855.stm.
51. “Professor Stephen Hawking and Virgin Galactic,” Virgin, December 17, 2015,
www.virgin.com/richard-branson/professor-stephen-hawking-and-virgin-galactic.
52. “Professor Stephen Hawking and Virgin Galactic.”
CHAPTER 6: BOUNDARIES
1. Author’s personal recollection of APS meeting in Atlanta, March 1999, written down twenty years
later.
2. “Hawking Draws Packed House to Atlanta Civic Center,” APS News 8, no. 5 (May 1999),
www.aps.org/publications/apsnews/199905/hawking.cfm.
3. “Remarks by President Clinton and Q&As at Hawking Lecture,” White House Millennium Council,
March 6, 2000, https://clintonwhitehouse4.archives.gov/Initiatives/Millennium/19980309-22774.html.
4. “Descent,” Star Trek: The Next Generation, Season 6, Episode 26, directed by Alexander Singer,
aired June 19, 1993.
5. The Simpsons, Season 10, Episode 22, directed by Pete Michels, aired May 9, 1999; Nicholas Hellen
and Steve Farrar, “Hawking Joins the Celestial High Earners,” Sunday Times (London), March 28,
1999, 3; “Specsavers—Stephen Hawking (1999),” YouTube, posted March 14, 2017, by The
Drum/Hosting, www.youtube.com/watch?v=aLMNYxEzXvU.
6. “Remarks by President Clinton and Q&As at Hawking Lecture.”
7. Matin Dunani and Peter Rodgers, “Physics: Past, Present, Future,” Physics World 12, no. 12
(December 1999): 7–14, https://doi.org/10.1088/2058-7058/12/12/2.
8. Errol Morris, personal communication with author.
9. Stephen Hawking, The Universe in a Nutshell (New York: Bantam Books, 2001), vii.
10. Kitty Ferguson and S. W. Hawking, Stephen Hawking: An Unfettered Mind (New York: St.
Martin’s Press, 2017), 199.
11. Neel Shearer, personal communication with author.
12. Jon Turney, “Review: The Universe in a Nutshell by Stephen Hawking,” The Guardian, November
10, 2001, www.theguardian.com/books/2001/nov/10/scienceandnature.highereducation.
13. “Rave Review for Hawking’s New Book,” Physics World, November 8, 2001,
https://physicsworld.com/a/rave-review-for-hawkings-new-book.
14. John Gribbin, “The Universe in a Nutshell by Stephen Hawking,” The Independent, November 3,
2001, www.independent.co.uk/arts-entertainment/books/reviews/the-univese-in-a-nutshell-by-sephen-
hawking-9153195.html.
15. Hawking, Universe in a Nutshell, 15.
16. “Hawking Rewrites History,” The Australian, October 15, 2001, 9.
17. Jane Hawking, Music to Move the Stars: A Life with Stephen (London: Pan, 2000), 573.
18. Jane Hawking, Music to Move the Stars, 580, 586.
19. “Stephen and Me: Libby Brooks Talks to Jane Hawking,” The Guardian, August 4, 1999, 1; Jane
Hawking, Music to Move the Stars, 555.
20. Tim Adams, “Jane Hawking: Brief History of a First Wife,” The Guardian, April 3, 2004; Jane
Hawking, Music to Move the Stars, 23–24, 46, 572.
21. Veronica Lee, “The Playwright, the Scientist, His Wife, and Her Lover,” Globe and Mail, August
15, 2000.
22. Robin Hawdon, God and Stephen Hawking, Josef Weinberger Plays, 2000.
23. Martin Anderson, “God and Stephen Hawking: The Playhouse, Derby, The Independent,”
September 5, 2000, www.independent.co.uk/arts-entertainment/theatre-dance/features/god-and-
stephen-hawking-the-playhouse-derby-701939.html; Matin Dunani, “Hawking Slams ‘Stupid, Worthless’
Play,” Physics World, August 2000, 8.
24. “All My Shootings Be Drivebys,” MC Hawking website, www.mchawking.com/all-my-shootings-
be-drivebys---lyrics.
25. Susan Carpenter, “Check It! MC Hawking Raps,” Los Angeles Times, November 2, 2000.
26. Roger Highfield, “I Thought My Time Was Up at the Age of 59.97, Says Hawking,” The
Telegraph, January 12, 2002, www.telegraph.co.uk/news/uknews/1381187/I-thought-my-time-was-up-
at-the-age-of-59.97-says-Hawking.html.
27. Hélène Mialet, Hawking Incorporated: Stephen Hawking and the Anthropology of the
Knowing Subject (Chicago: University of Chicago Press, 2012), 2.
28. Neil Turok, personal communication with author.
29. Peter Stringfellow, personal communication with author, July 2003.
30. “Stephen Hawking at 70: Exclusive Interview,” New Scientist, January 4, 2012,
www.newscientist.com/article/mg21328460-500-stephen-hawking-at-70-exclusive-interview.
31. Hawking, Brief History, 136.
32. Don N. Page, “The Hartle-Hawking Proposal for the Quantum State of the Universe,” in The
Creation of Ideas in Physics: Studies for a Methodology of Theory Construction, ed. Jarrett
Leplin, University of Western Ontario Series in Philosophy of Science (Dordrecht: Springer Netherlands,
1995), 184.
33. S. W. Hawking, “The Cosmological Constant Is Probably Zero,” Physics Letters B 134, no. 6
(January 26, 1984): 403.
34. Neil Turok, personal communication with author.
35. Kip S. Thorne, Black Holes and Time Warps: Einstein’s Outrageous Legacy (New York: Norton,
1994), 419; Andy Strominger, personal communication with author.
36. “The Hawking Paradox, Horizon, Season 42, Episode 2, directed by William Hicklin, BBC, aired
September 15, 2005.
37. Leonard Susskind, The Black Hole War: My Battle with Stephen Hawking to Make the World
Safe for Quantum Mechanics (Boston: Little, Brown, 2009), 419.
38. Peter D’Eath, personal communication with author (via email) for this quotation and others that
immediately follow.
39. Andrew Farley, personal communication with author (via email).
40. Christophe Galfard, personal communication with author; “The Hawking Paradox”; Mialet,
Hawking Incorporated, 93.
41. “The Hawking Paradox,” Season 42, Episode 2.
42. Pallab Ghosh, “The Day I Thought We’d Unplugged Stephen Hawking,” BBC, March 14, 2018,
www.bbc.com/news/science-environment-43400021.
43. Neil Turok, personal communication with author.
CHAPTER 7: INFORMATION
1. “A Brief History of Physicist Hawking’s Plans to Remarry,” Chicago Tribune, July 6, 1995, 2.
2. Jane Hawking, Music to Move the Stars: A Life with Stephen (London: Pan, 2000), 562, 580; Ray
Laflamme, personal communication with author.
3. Jane Hawking, Music to Move the Stars, 585; Matrimonial Causes Act of 1973, 28(3), rev. February
1, 1991.
4. Matrimonial Causes Act of 1973, 1(2), rev. February 1, 1991.
5. “Hawking to Marry His Former Nurse,” The Times (London), July 6, 1995; Jane Hawking, Music to
Move the Stars, 585; “Hawking Gets Hitched,” The Gazette, September 16, 1995. Leonard Mlodinow,
in his Stephen Hawking: A Memoir of Friendship and Physics (New York: Pantheon Books, 2020),
214, states that Robert, the eldest child, was, in fact, in attendance, but that the other two were not.
6. Personal communications with author.
7. For more about information and information theory, see my Decoding the Universe: How the New
Science of Information Is Explaining Everything in the Cosmos, from Our Brains to Black Holes
(New York: Viking, 2006).
8. Leonard Susskind, The Black Hole War: My Battle with Stephen Hawking to Make the World
Safe for Quantum Mechanics (Boston: Little, Brown, 2009), 184.
9. “1990s,” Royal Albert Hall, www.royalalberthall.com/about-the-hall/our-history/explore-our-
history/time-machine/1990s.
10. “Show Detail: Aspen Center for Physics Presents: Does God Throw Dice in Black Holes with
Stephen Hawking, July 24, 1996,” Video On-Demand, Cablecast,
http://mc.grassrootstv.org/Cablecast/public/Show.aspx?ChannelID=1&ShowID=947. There might be
slight differences between this version of the lecture and the one given several months earlier at Albert
Hall.
11. “A Matter of Divine Right: Does God Play Dice with the Universe? Einstein Said No, but Stephen
Hawking Says Yes. Peter Millar Reports,” South China Morning Post, December 4, 1995, 20.
12. Stephen Hawking, Black Holes and Baby Universes and Other Essays (New York: Bantam
Books, 1993), 49 ff.
13. Hawking, Black Holes and Baby Universes, 67.
14. Daniel Z. Freedman, personal communication with author.
15. Sara Lippincott, “Interview with John H. Schwarz,” 2002, Caltech Oral Histories, Caltech Institute
Archives, https://resolver.caltech.edu/CaltechOH:OH_Schwarz_J, 44.
16. Stephen Hawking, typescript [photocopy] of a draft of A Brief History of Time with typed letter
signed from Hawking to Mr. S. [Stephen] Zatman, March 24, 1986, Washington University, St. Louis,
ID no. MSS/VMF/134, p. 94/101; Stephen Hawking, A Brief History of Time (New York: Bantam
Books, 1988), 165.
17. Marika Taylor, personal communication with author.
18. Laflamme, personal communication with author.
19. Susskind, Black Hole War, 391.
20. Andy Strominger, personal communication with author.
21. Susskind, Black Hole War, 394.
22. Taylor, personal communication with author for this quotation and others that immediately follow.
23. Valerie Grove, “The Million-Dollar Cinderella,” The Times (London), February 7, 1997, 16.
24. Daya Alberge, “Tolkien Wins Title Lord of the Books by Popular Acclaim,” The Times (London),
January 20, 1997.
25. Robert Crampton, “Intelligence Test,” The Times (London), April 8, 1995.
26. Al Zuckerman, personal communication with author. The elision excises Zuckerman’s mistaken
reference to Universe in a Nutshell (also packaged by Philip Dunn) rather than The Illustrated Brief
History of Time, which was the subject being discussed.
27. “The Big Bang,” Stephen Hawking’s Universe, Episode 6, aired on PBS in 1997.
28. “US Robotics X2 Commercial (Stephen Hawking),” YouTube, posted May 18, 2012, by vector108,
www.youtube.com/watch?v=-HAUiXRuH_I; “Physicist Hawking Getting a Big Bang Out of Internet,”
Chicago Tribune (Associated Press), March 23, 1997, 19.
29. Paul Rodgers, “$6m Hawking Professorship Is ‘Too Generous,’ Say Critics,” Forbes, February 5,
2014, www.forbes.com/sites/paulrodgers/2014/02/05/us-philanthropists-6m-for-stephen-hawking-
professorship-is-too-generous-say-cambridge-critics; Neil Turok, personal communication with author.
30. Michael D. Lemonick, “Hawking: Is He All He’s Cracked Up to Be?,” Time, February 3, 2014,
https://time.com/3531/hawking-myth-or-legend.
31. Peter Kingston, “Curtain Up on the Cosmos,” The Guardian, November 13, 1997; Falling
Through a Hole in the Air, written by Judith Goldhaber and Carlton Reese Pennypacker, directed by
David Parr, 2007, recording provided by Carlton Reese Pennypacker.
32. Hélène Mialet, Hawking Incorporated: Stephen Hawking and the Anthropology of the
Knowing Subject (Chicago: University of Chicago Press, 2012), 58.
33. Mialet, Hawking Incorporated, 61.
34. Wager posted at Preskill’s Caltech website. See
www.theory.caltech.edu/people/preskill/old_naked_bet.html.
35. John Preskill, personal communication with author.
36. Kip Thorne, “Warping Spacetime,” in The Future of Theoretical Physics and Cosmology:
Celebrating Stephen Hawking’s 60th Birthday, ed. G. W. Gibbons, S. J. Rankin, and E. P. S. Shellard
(Cambridge: Cambridge University Press, 2003), 24.
37. Preskill, personal communication with author.
38. Wager posted at Preskill’s Caltech website. See
www.theory.caltech.edu/people/preskill/info_bet.html.
39. Susskind, Black Hole War, 419.
40. Strominger, personal communication with author.
41. Joe Friesen, “Hawking Proves a Good Sport When It Comes to Settling Bets,” Globe and Mail,
July 22, 2004.
42. Marika Taylor, personal communication with author.
CHAPTER 8: IMAGES
1. Emma Wilkins, “Wheelchair Physicist in Love Tangle with Nurse,” Daily Mail, August 1, 1990.
2. “People,” United Press International, July 29, 1990, BC cycle; Wilkins, “Wheelchair Physicist.”
3. Jane Hawking, Travelling to Infinity: My Life with Stephen Hawking (Richmond, UK: Alma
Books, 2014), 473.
4. Wilkins, “Wheelchair Physicist.”
5. Peter Guzzardi, personal communication with author.
6. Jane Hawking, Music to Move the Stars: A Life with Stephen (London: Pan, 2000), 444.
7. Hawking, Music to Move the Stars.
8. Bob Sipchen, “Simply Human: Profile. Wheelchair-Bound Physicist Stephen Hawking Resists Efforts
to Deify His Life or His Disabilities,” Los Angeles Times, June 6, 1990, E1.
9. Stephen Hawking, Black Holes and Baby Universes and Other Essays (New York: Bantam
Books, 1993), 32; Stephen Hawking, My Brief History (London: Bantam Books, 2018), 122; Hélène
Mialet, Hawking Incorporated: Stephen Hawking and the Anthropology of the Knowing Subject
(Chicago: University of Chicago Press, 2012), 71.
10. Leslie Hanscom, “I Have No Enemies,” Sydney Morning Herald, July 2, 1988.
11. Kip S. Thorne, Black Holes and Time Warps: Einstein’s Outrageous Legacy (New York: Norton,
1994), 413.
12. Michael S. Morris and Kip S. Thorne, “Wormholes in Spacetime and Their Use in Interstellar
Travel: A Tool for Teaching General Relativity,” American Journal of Physics 56, no. 5 (May 1988):
397.
13. “Citizen Science,” 365 Days of Astronomy, December 27, 2012,
https://cosmoquest.org/x/365daysofastronomy.
14. Morris and Thorne, “Wormholes in Spacetime and Their Use,” 407.
15. S. W. Hawking, “Chronology Protection Conjecture,” Physical Review D 46, no. 2 (July 15, 1987):
603.
16. Hawking, “Chronology Protection Conjecture.”.
17. Hawking, “Chronology Protection Conjecture.”
18. Hawking, “Chronology Protection Conjecture.”
19. Thorne, Black Holes and Time Warps, 521.
20. Sue Lawley, “Desert Island Discs: Stephen Hawking,” BBC Radio 4, December 27, 1992,
www.bbc.co.uk/programmes/p0093xb2.
21. John Preskill, personal communication with author.
22. Kitty Ferguson and S. W. Hawking, Stephen Hawking: An Unfettered Mind (New York: St.
Martin’s Press, 2017), 168; Judy Bachrach, “A Beautiful Mind, an Ugly Possibility,” Vanity Fair,
January 25, 2004, www.vanityfair.com/news/2004/06/hawking200406.
23. Morgan Strong, “Playboy Interview: Stephen Hawking,” Playboy, April 1990, 74.
24. Strong, “Playboy Interview,” 68, 64.
25. Bryan Appleyard, “A Master of the Universe,” Sunday Times (London), June 19, 1988, 30; Jane
Hawking, Music to Move the Stars, 592.
26. Lawley, “Desert Island Discs.”
27. “1993 British Telecom—‘Hawking’ TV Commercial,” YouTube, posted March 14, 2018, by History
of Advertising Trust, www.youtube.com/watch?v=GH5Q54eIaPk.
28. Sean Michaels, “Stephen Hawking Sampled on Pink Floyd’s The Endless River,” The Guardian,
October 8, 2014, www.theguardian.com/music/2014/oct/08/stephen-hawking-sampled-pink-floyd-the-
endless-river; “Radio Interview on Opening Night of 1994 Tour: The Division Bell,” Pink Floyd & Co.,
https://pfco.neptunepinkfloyd.co.uk/band/interviews/grp/grpredbeard.html.
29. Don N. Page and Stephen W. Hawking, “Gamma Rays from Primordial Black Holes,”
Astrophysical Journal 206 (May 15, 1976): 2.
30. Most sources give the date of this quotation as Hawking’s 2016 Reith Lecture, but it dates back at
least several years before that. S. W. Hawking, “Into a Black Hole,” lecture ca. 2008, Internet Archive,
Wayback Machine, https://web.archive.org/web/20120112012649/http://www.hawking.org.uk/into-a-
black-hole.html.
31. S. W. Hawking and J. M. Stewart, “Naked and Thunderbolt Singularities in Black Hole
Evaporation,” arxiv:hep-th/9207105v1, July 30, 1992.
32. Raphael Bousso, personal communication with author.
33. Mialet, Hawking Incorporated, 55; Roger Penrose, “‘Mind over Matter’: Stephen Hawking—
Obituary by Roger Penrose,” The Guardian, March 14, 2018,
www.theguardian.com/science/2018/mar/14/stephen-hawking-obituary.
34. R. Bousso and S. W. Hawking, “Black Holes in Inflation,” Nuclear Physics B 57 (1997): 204.
35. Bousso, personal communication with author.
36. Hawking, “Into a Black Hole.”
37. Leonard Susskind, “Lecture 8: Black Hole Formation, Penrose Diagrams and Wormholes.
CosmoLearning Physics,” CosmoLearning, posted January 11, 2015, https://cosmolearning.org/video-
lectures/black-hole-formation-penrose-diagrams-wormholes; Mialet, Hawking Incorporated, 68–70.
38. Mialet, Hawking Incorporated, 68–70.
39. Roger Penrose, personal communication with author.
40. G. H. Hardy, A Mathematician’s Apology (Cambridge: Cambridge University Press, 1940), 6–7.
41. Andreas Albrecht, personal communication with author.
42. Robert Crampton, “Intelligence Test,” The Times (London), April 8, 1995.
43. Graham Farmelo, The Strangest Man: The Hidden Life of Paul Dirac, Mystic of the Atom (New
York: Basic Books, 2009), 288.
44. Hanscom, “I Have No Enemies.”
45. Hanscom, “I Have No Enemies.”
46. Hawking, A Brief History of Time, 116.
47. Hawking, A Brief History of Time, 174.
48. William Craig, “‘What Place, Then, for a Creator?’: Hawking on God and Creation,” British
Journal for the Philosophy of Science 41, no. 4 (December 1990): 483.
49. Hawking, Black Holes and Baby Universes, 41–42.
50. Jane Hawking, Music to Move the Stars, 536. The other journalist was Pauline Hunt, who wrote
for the local Cambridge paper.
51. “Master of a Narrow Universe: Stephen Hawking Is on a Voyage to Stardom but Unable to
Navigate in the Human Realm,” The Independent, October 13, 1993,
www.independent.co.uk/voices/master-of-a-narrow-universe-stephen-hawking-is-on-a-voyage-to-
stardom-but-unable-to-navigate-in-the-1510340.html.
52. Robert Crampton, “Intelligence Test,” The Times (London), April 8, 1995.
53. Martyn Harris, “A Brief History of Hawking,” Spectator Archive, June 27, 1992,
http://archive.spectator.co.uk/article/27th-june-1992/18/a-brief-history-of-hawking.
CHAPTER 9: FLASH
1. Ray Laflamme, personal communication with author.
2. Peter Guzzardi, personal communication with author.
3. Stephen Hawking, typescript [photocopy] of a draft of A Brief History of Time with typed letter
signed from Hawking to Mr. S. [Stephen] Zatman, March 24, 1986, Washington University, St. Louis,
ID no. MSS/VMF/134 (also available at the University of Cambridge Library, Add. MS 9222, 95).
4. Stephen Hawking, “Is the End in Sight for Theoretical Physics?,” in Black Holes and Baby
Universes and Other Essays (New York: Bantam Books, 1993), 65.
5. Al Zuckerman, personal communication with author; John Gribbin, personal communication with
author.
6. Guzzardi, personal communication with author for this quotation and others that immediately follow.
7. Archana Masih, “The Stephen Hawking I Knew,” Rediff, March 22, 2018,
www.rediff.com/news/special/the-stephen-hawking-i-knew/20180322.htm.
8. Hawking, Brief History typescript, 3.
9. Stephen Hawking, “Drafts of Stephen Hawking’s A Brief History of Time,” ca. 1983–ca. 1987,
University of Cambridge Library, Add. MS. 9222, 171.
10. Stephen Hawking, A Brief History of Time (New York: Bantam Books, 1998), 1.
11. Hawking, Brief History typescript, 95.
12. Hawking, A Brief History of Time, 175.
13. Peter Guzzardi, personal communication with author; Stephen Hawking, “A Brief History of A Brief
History,” Popular Science, August 1989, 70 ff.
14. Masih, “The Stephen Hawking I Knew.”
15. Guzzardi, personal communication with author.
16. Don Page, “Hawking’s Timely Story,” Nature 332 (April 21, 1988): 742–743; Hawking, A Brief
History of Time, 73, 95.
17. Hawking, “A Brief History of a Brief History”; Elizabeth Mehren, “Reagan Administration Book
Beat Picks Up,” Los Angeles Times, April 3, 1988, www.latimes.com/archives/la-xpm-1988-04-03-bk-
1130-story.html; Michael White and John Gribbin, Stephen Hawking: A Life in Science (New York:
Dutton, 1992), 286.
18. Jane Hawking, Music to Move the Stars: A Life with Stephen (London: Pan, 2000), 486.
19. Andy Albrecht, personal communication with author.
20. White and Gribbin, Stephen Hawking, 326, 327.
21. Hawking, “Drafts of Stephen Hawking’s A Brief History of Time,” ca. 1983–ca. 1987, 282; Stephen
Hawking, A Brief History of Time, 131.
22. White and Gribbin, Stephen Hawking, 327.
23. Albrecht, personal communication with author; White and Gribbin, Stephen Hawking, 328; Jerry
Adler, “Reading God’s Mind,” Newsweek, June 13, 1988.
24. Dennis Overbye, Lonely Hearts of the Cosmos (New York: Harper and Row, 1991), 253.
25. Stephen Hawking, “Inflation Reputation Reparation,” Physics Today 42, no. 2 (February 1989): 16.
26. Albrecht, personal communication with author.
27. Stephen Hawking, “Why Does the Universe Inflate?,” in Quantum Mechanics of Fundamental
Systems: The Quest for Beauty and Simplicity, ed. Marc Henneaux and Jorge Zanelli (New York:
Springer Science + Business Media, 2009), 147; Daniel Z. Freedman, personal communication with
author.
28. Adler, “Reading God’s Mind.”
29. “Suiting Science to a T (Shirt), Two Chicago Bar Owners Set Up a Stephen Hawking Fan Club,”
People, September 11, 1989, https://people.com/archive/suiting-science-to-a-t-shirt-two-chicago-bar-
owners-set-up-a-stephen-hawking-fan-club-vol-32-no-11; Steven Pratt, “His Fan Club Suits Hawking to
a T,” Chicago Tribune, July 26, 1989, www.chicagotribune.com/news/ct-xpm-1989-07-26-8902200569-
story.html.
30. “Suiting Science to a T.”
31. Laflamme, personal communication with author.
32. Nigel Farndale, “A Brief History of the Future,” The Independent, January 8, 2000,
www.independent.ie/irish-news/a-brief-history-of-the-future-26125776.html.
33. Laflamme, personal communication with author; Jane Hawking, Music to Move the Stars, 501.
34. Charles Oulton, “Cosmic Writer Shames Book World,” Sunday Times (London), August 28, 1988.
35. Hawking, “A Brief History of a Brief History,” 70 ff.
36. David Blum, “The Tome Machine,” New York Magazine, October 24, 1988, 40.
37. Hawking, “A Brief History of a Brief History.”
38. Guzzardi, personal communication with author.
39. Simon Mitton, personal communication with author.
40. Morgan Strong, “Playboy Interview: Stephen Hawking,” Playboy, April 1990, 64.
41. Hawking, Black Holes and Baby Universes, 1.
42. Hawking, My Brief History, 6.
43. Errol Morris, personal communication with author.
44. Morris, personal communication with author.
45. Morris, personal communication with author.
46. Morris, personal communication with author.
47. Morris, personal communication with author.
48. Raphael Bousso, personal communication with author.
49. Jane Hawking, Music to Move the Stars, 504, 505.
50. Hawking, My Brief History, 85; Jane Hawking, Music to Move the Stars, 339.
51. Jane Hawking, Music to Move the Stars, 455, 460.
52. Jane Hawking, Music to Move the Stars, 487, 538.
53. Jane Hawking, Music to Move the Stars, 540.
54. Jane Hawking, Music to Move the Stars, 548.
55. Jane Hawking, Music to Move the Stars, 546.
56. Jane Hawking, Music to Move the Stars, 553.
57. Richard Jerome, “Of a Mind to Marry,” People, August 7, 1995, https://people.com/archive/of-a-
mind-to-marry-vol-44-no-6.
58. Jane Hawking, Music to Move the Stars, 557.
59. Bryan Appleyard, “Stephen Hawking Changed My Life,” Sunday Times (London), December 14,
2014.
60. Pliny, Natural History, Book 33, Chapter 4; Tertullian, Apology, 33.3.
61. Bryan Appleyard, “A Master of the Universe,” Sunday Times (London), June 19, 1988, 30.