Fear of A Black Universe - Stephon Alexander

Download as pdf or txt
Download as pdf or txt
You are on page 1of 221
At a glance
Powered by AI
The document provides an overview of a physics book that discusses mysterious phenomena like dark matter and dark energy, as well as topics like the origin of the universe and life in the cosmos.

The document is structured with page numbers and text from different parts of the book, including the table of contents, introductions of parts, chapters, and praise for the book.

The book covers topics in physics like cosmology, quantum fields, emergence, as well as cosmic phenomena like the Big Bang, dark matter, and the possibility of alien life.

Copyright © 2021 by Stephon Alexander

Cover design by Ann Kirchner


Cover image © Chinnapong / Shutterstock.com
Cover copyright © 2021 by Hachette Book Group, Inc.

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.

The scanning, uploading, and distribution of this book without


permission is a theft of the author’s intellectual property. If you would
like permission to use material from the book (other than for review
purposes), please contact [email protected]. Thank you for
your support of the author’s rights.

Basic Books
Hachette Book Group
1290 Avenue of the Americas, New York, NY 10104
www.basicbooks.com

First Edition: August 2021

Published by Basic Books, an imprint of Perseus Books, LLC, a


subsidiary of Hachette Book Group, Inc. The Basic Books name and
logo is a trademark of the Hachette Book Group.

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.

Images © Stephon Alexander except Figure 26 © Volker Springel


and Figure 27 © Jiro Soda
Library of Congress Cataloging-in-Publication Data
Names: Alexander, Stephon, author.
Title: Fear of a black universe : an outsider’s guide to the future of
physics / Stephon Alexander.
Description: First edition. | New York : Basic Books, 2021. | Includes
bibliographical references and index.
Identifiers: LCCN 2021001977 | ISBN 9781541699632 (hardcover) |
ISBN 9781541699618 (ebook)
Subjects: LCSH: Cosmology. | Physics—Research—Methodology. |
Research—Philosophy. | Research—Social aspects.
Classification: LCC QB981 .A538 2021 | DDC 523.1—dc23
LC record available at https://lccn.loc.gov/2021001977

ISBNs: 978-1-5416-9963-2 (Hardcover); 978-1-5416-9961-8 (Ebook)

E3-20210723-JV-NF-ORI
CONTENTS

Cover
Title Page
Copyright
Dedication

PART I
1 Escape from the Jungle of No Imagination
2 The Changeless Change
3 Superposition
4 The Zen of Quantum Fields
5 Emergence
6 If Basquiat Were a Physicist

PART II
Cosmic Improvisations
7 What Banged?
8 A Dark Conductor of Quantum Galaxies
9 Cosmic Virtual Reality
10 Embracing Instabilities
11 A Cosmologist’s View of a Quantum Elephant
12 The Cosmic Biosphere
13 Dark Ideas on Alien Life
14 Into the Cosmic Matrix
15 The Cosmic Mind and Quantum Cosmology

Acknowledgments
Discover More
About the Author
Notes
Also by Stephon Alexander
Praise for Fear of a Black Universe
In loving memory of my grandmothers, Celisha Belfon and Ruby
Farley, who taught me how to heal with the imagination.
Explore book giveaways, sneak peeks, deals, and more.

Tap here to learn more.


PART I
1

ESCAPE FROM THE JUNGLE OF


NO IMAGINATION

Into blinding darkness enter


Those who worship ignorance
Into as if still greater darkness
Enter those who delight in knowledge
—THE UPANISHADS

We physicists have determined that over 95 percent of the matter


and energy in the universe is invisible. We have branded this
enigmatic stuff dark matter and dark energy; their discovery raised
puzzles that shook the foundations of physical law. The gravitational
effects of dark matter are observed in large halos surrounding
galaxies and are critical to our current conception of how the large-
scale arrangement of visible matter in the universe came to be.
Likewise, so far, with dark energy, which was discovered with
telescopes by measuring the accelerated expansion of the universe,
it too has been the province of cosmologists, who have written about
it only in reference to extraterrestrial matters and the overall shape
and destiny of the universe. This is a mistake. This dark stuff turns
out to play a hidden role in the visible world, including in our
understanding of life itself. Dark energy resides in all empty space,
not just outer space, and permeates all existence. Its quantum
effects are present even in the spaces between the very atoms in
our bodies. The time has come for a new Newton, to reunite the
physics of the extraterrestrial with the physics of the terrestrial. Such
an integration might facilitate our understanding of dark matter and
dark energy, enabling a better understanding of who we are and of
the cosmos in which we live.
Just as the discoveries of dark matter and dark energy shook the
foundations of physics, our continued inability to unearth the identity
and nature of most of the universe continues to shake them, and,
consequently, it limits our understanding of our place in the universe.
We still do not know much about the dark sector except that it exists;
yet researchers often ascribe properties to dark matter based on
presumptions that mimic known physics and are not intrepid enough.
It seems to me that methodologies that might enable us to ask new
questions, and find new properties or new roles for the dark in our
universe, generate fear. Do we dread the dark so much that we
project our fears onto the very phenomena about which we are
scientifically ignorant?
Dark matter and dark energy are not the only anomalies our
current physics doesn’t handle. A handful of other deviations from
our accepted theories of physics generate speculations that likewise
trouble physics. The resolution of these anomalies may shake the
foundation of what we presume to be true.
Such anomalies raise a related set of questions, one more
apropos for the social sciences than the natural sciences: How does
science respond to ideas that might violate our scientific norms and
expectations? Does the scientific community fear embracing “dark”
ideas from outsiders, especially if the ideas may not be in a form that
the community is comfortable with, if they do not fit seamlessly into
our theories and expected practices?
At the turn of the last century, the discovery of black-body
radiation found in most objects that appeared to not emit light was a
theoretical “catastrophe” for classical physics, giving nonsensical
predictions that are not seen in experiments. But when German
physicist Max Planck embraced the reality of the black body, he
turned electromagnetic theory on its head, and the quantum
revolution was born. Is it possible that the theoretical anomalies we
now confront will yield a comparable scientific revolution? If so, who
is likely to motivate it?
Regardless of our ability to create the most abstract mathematics
and come to know truths beyond our five senses, as humans we are
limited by our social and psychological conditioning. In this book, we
will go beyond the current conceptual and scientific-sociological
paradigm into uncharted and sometimes risky conceptual territories.
What lurks beyond the black hole singularity in our galaxy and before
time existed at the big bang? How did cosmic structure emerge from
a chaotic and featureless evolving early universe? What is the role of
dark energy and dark matter in the universe? Is there a hidden link
between the emergence of life and the laws of physics? These are
questions on the boundary of what we know; answering them may
call into question the theories that constitute our knowledge. If we
are to answer them, we must ask whether the scientific community is
able to incorporate into its activities nontraditional members,
outsiders more likely to see beyond our current theoretical horizon;
further, is the scientific community, as it is now structured, able to
empower these outsiders to break new ground?
I want to bring you with me as I try to take on some of these
questions. To do this effectively, I will provide both the necessary
background and the conceptual tools needed to understand a bit of
established physics. My discussion is based on three fundamental
principles that underlie all known physics; a grasp of them will enable
us to understand some of the problems at the borders of what
physicists think they know and understand. I will be frank, sometimes
controversial, and deliberately engage in some of my own wild
speculations. This is not just a book about what we know in physics,
but a book that explicates the frontiers of physics, a book about how
physics is done.
Much that is taught and written about physics expresses what we
know already. This book presumes that the process of doing physics
is different than the process of learning the physics we know already.
The first explores what we do not yet know, while the second
transmits what others have learned previously. Here, while some of
the latter is necessary—there are certain things that must be shared
—our focus is on how we might think about what we do not yet know.
Crucial is a simple insight: a responsibility of physicists is to apply
what we know already in new areas of inquiry, to transform and
extend our knowledge. Great teaching in physics helps us to do
physics, not simply learn the physics we know already. This means
learning physics has to enable us to work at the boundary of what
we know or, in rare cases, to go beyond those boundaries, or even
reconstruct the very framework of our knowledge. If this book is
successful, it will help you understand what it means to be creative in
theoretical physics.

Often when I am stuck on a problem—of the physics variety or the


personal one—I make a pilgrimage to the northern coast of my
birthplace, Trinidad and Tobago. There is something that feels
unspeakably out of body about trekking through the lush sixty-mile
stretch of the deep green mountain range overlooking Las Cuevas
Bay. I hike up the winding paths to the top of a hill overlooking the
ocean, the tropical jungle sounds looming behind me and the
rhythmic crashing of the crystalline emerald crescent waves
sounding below. Surrounded by nature, beautiful and primordial, I
am often surprised to find new insights into my problems.
One day not so long ago, I found myself getting nowhere on a
research problem. I headed back to the jungle to look at the sea.
While I was there it dawned on me—not the solution to the research
problem, but the realization that during two decades of scientific
research, I had been unconsciously dodging my original reason for
becoming a physicist: to make a meaningful scientific discovery. I
realized I feared failure and the professional risk failure entailed. The
ability to maintain a scientific career is driven by, among other
factors, your reputation among your peers and familiarity with your
work. Penalties await those who are perceived as a “crackpot” or
who speculate too much. I knew that some of the ideas that
interested me, such as the connection between consciousness and
quantum mechanics, would make me vulnerable to stigma and
potentially stump my career.
In theoretical physics research, there is a sense of dissatisfaction,
a belief that we have not been able to break new ground in the same
way that led to the quantum and relativity revolutions early last
century. It’s not to say that people aren’t trying to address their
dissatisfactions; a handful of papers are posted every day on an
online global archive of physics research called the Archives, and
oftentimes these papers offer new approaches to unsolved
mysteries. Despite this, there’s not much feeling of progress. Why is
this? Is it because these problems are too hard for us? Or is it that in
the search for the truth, some scientists are afraid to look at
uncharted or forbidden territories, afraid because there may be
penalties, reputational and professional, for stepping outside
accepted paradigms? I think that it’s the latter. In this book, I will
provide my thoughts and reflections; I will take some risks, hoping
that we learn something significant along the way, whether I am right
or wrong.
As a Black physicist, this potential strength—that I am brimming
with ideas, my capacity to generate speculative thinking—can be an
impediment. Black persons in scientific circles are often met with
skepticism about their intellectual capabilities, their ability to “think
like a physicist.” Consequently, my exploratory, personal style of
theorizing, when coupled with my race, often creates situations
where my white colleagues become suspicious and devalue my
speculations. I have navigated a career in physics in spite of these
racial and sociological prejudices, and, given both my personality
and my predilections, I continue to march ahead, sharing my
conjectures, which, at least sometimes, are theoretically fruitful. This
book will be no exception.
During my time of self-reckoning in Trinidad, I decided to devote
the majority of my research efforts to working on some of the big
mysteries in physics. To do so effectively, I would have to bring my
entire being to how I do physics, which meant engaging in
improvisational and wild speculations. When you meet me in person,
it is clear that I am volcanic with ideas, most of which turn out to be
wrong, while some, even among those that are “wrong,” are fruitful
and worth pursuing. Underlying these ideas is a latent foundation,
the theoretical and technical tools of my trade.
Physics is a social activity, and like all social activities it is
regulated by norms. Practitioners are expected to conform to these
normative expectations, and they are sanctioned negatively when
they violate them. Too often the expectations of what it means to do
“good science” become confused with specific theoretical
orientations, which means that practitioners in subdisciplines are
expected to uphold specific theoretical arguments. This is desirable
insofar as it rules out ideas like flat-earthism and others that make no
sense scientifically. Sometimes, however, this expectation of
conformity stifles innovation and progress. Some scientists are
reluctant to explore ideas outside the expected paradigm because
they will be punished if they do so, which means that paradigm-
shattering theories can be inhibited from emerging.
We need to distinguish clearly between the values and norms that
regulate scientific activity and those that demand conformity to a
particular body of theory, a particular paradigm, within a scientific
community. Both are constituted socially, but the latter obligations
can restrict our creativity, our ability to constitute new theoretical
orientations. It is crucial, however, to recognize that our theoretical
arguments must be regulated within and evaluated through the
application of scientific values, the values of cognitive rationality.
Very simply, this means that our theoretical arguments must be
logically coherent and empirically warrantable. Not every “creative
idea” may be turned into viable physical theory. In fact, the likelihood
that any one of us will create a new paradigm because we have
violated the norms regulating activity within the standard paradigm is
very slim. No one can do so, however, without violating these norms.
I want this book to serve as a source of inspiration and
encouragement for individuals who feel disenfranchised and
unwelcome in our scientific communities, people who are
sometimes, or often, made to feel that they are not valued as
contributors to the scientific endeavor. So as much as this is a book
about my reflections on the state of physics, as theory, I also reflect
on and analyze both the sociology of science and my own
experiences to argue for the efficacy of outsiders’ presence and
perspectives in scientific communities and inquiry. The path to
becoming a scientist poses challenges for everyone. In shedding
new light on the social dynamics of science, and simply sharing our
stories, we can see how some of the challenges outsiders face can
inspire them to make significant scientific contributions. I hope to
convince my readers that diversity in science is not simply a social
justice concern, but that it enhances the quality of the science we
accomplish.

Many of the theoretical physicists of my generation were inspired by


the golden era in the first half of the twentieth century, when the likes
of Albert Einstein, Richard Feynman, Paul Dirac, Emmy Noether,
and Wolfgang Pauli, to name only a few of our idols, gave birth to
quantum field theory and general relativity. These theories have
been spectacularly confirmed, and they are responsible for most of
our modern technology.
One of the essential tools that Einstein and Erwin Schrödinger
employed in discovering the equations and fundamental laws of
relativity and quantum physics was “thought experiments”: mental
visualizations, or imaginations of physical happenings that are
impossible to carry out in terrestrial settings or with current
experimental techniques. Some of the famous ones include
Schrödinger’s cat and Einstein’s vision of riding on a beam of light.
These visualizations, when articulated as mathematical equations,
led to solutions that predict the behavior of the semiconductor
devices that drive powerful computers, including the smartphones
that are part of our everyday lives.
When I first learned how the greats managed to make these
discoveries, it seemed as if some mental wizardry were at work, a
wizardry that has been overlooked by my generation and our
teachers. Theoretical physics has grown to become extremely
mathematical, and while mathematics is a necessary and powerful
tool, I realized that if I were to have a shot at making an important
discovery, I would have to find my way to acquire a bit of that
wizardry, the intuitions leading to the theoretical insights that lead to
mathematical equations (intuitions not derivative from those
equations).
As a young student taking introductory courses in physics, I had
the impression that physics was a jungle of countless equations and
intricate theories. The task, or so it seemed, was to digest and apply
them. Even decades later, as a researcher in theoretical physics, it
dawned on me that my colleagues and I were lost in that same
jungle. The mentality required to work through problem sets made
the handful of problems in cosmology and particle physics that
seemed important also seem insurmountable. We did not even know
the right questions to ask.
At Las Cuevas Bay, after gazing at the waves for some time, I
had an epiphany: Who better to help us address our questions than
Einstein himself? What if we had a bird’s-eye view of the jungle of
physics from which we could see the origins of the theories and the
interconnections between the laws that give rise to (and constrain)
them? Would this perspective facilitate our attempts at reworking
these theories to better address our contemporary questions? Could
we turn from calculating, boring physicists to brave adventurers,
imagining worlds no one else had seen before?
During my time as a postdoctoral researcher in theoretical
physics at the Stanford Linear Accelerator Center (SLAC), I received
a surprising letter from the National Geographic Society. I wondered
if I owed them a payment. Instead, the letter congratulated me on
being selected as a National Geographic Emerging Explorer. I was
both elated and confused. I had never applied to be an explorer, nor
did I think of myself as one. When it turned out that it was not, in fact,
a mistake, I was deeply honored and did not say no to the monetary
prize and subsequent trips to National Geographic headquarters to
meet other explorers I admired. For example, I had always wanted to
meet ethnobotanist Wade Davis, whose book was the basis for one
of my favorite horror movies, The Serpent and the Rainbow.
All explorers were invited to a fundraiser and to celebrate the
seventieth birthday of the Society’s president at the time, Gil
Grosvenor. There were many impressive people there, and I quickly
started to feel a little out of place. Among the newly elected explorers
was an underwater cave diver who could contort his body to fit into
intricate caves for hours, hundreds of feet under the ocean. There
was a woman who lived among lions in the Serengeti, and a man
who explored and lived in Antarctica for extended periods. At the
fundraiser, each explorer was placed at a dinner table with a group
of potential donors, to entertain them. After we introduced ourselves,
one disappointed donor said to me, “You don’t hang out in the
jungle? You don’t fly airplanes? Why did they make a theoretical
physicist an explorer?!”
I didn’t want the donor to feel duped. So as a good spokesperson
for National Geographic, I responded with conviction: “I explore the
cosmos with my mind.” I went on to explain how the worlds that
cosmologists explore are even more extreme than explorers on
Earth, so extreme that we are forced to explore them in our
imaginations. I went on to explain that Einstein explored the nature
of space-time, and this led to the ultimate prediction and discovery of
a supermassive black hole at the center of our galaxy. Try exploring
that physically! Some donors were interested, but others wanted to
hang out with a “real” explorer.
Despite the drama, that night got me thinking about the
similarities between physical and mental exploration, about the
extreme places theoretical physicists must explore to make
progress. These mental explorations are the fuel for discovering and
clarifying physical theories; they are the domain of Einstein’s notion
of principle theories.
In 1914, soon after his revolutionary discoveries in quantum
mechanics and relativity, Einstein gave an address to the Prussian
Academy of Sciences in which he discussed his strategy for
discovery in theoretical physics. “The theorist’s method involves his
using as his foundation general principles from which he can deduce
conclusions,” Einstein said. “His work thus falls into two parts. He
must first discover his principles and then draw the conclusions
which follow from them.”
Einstein could perceive a hidden reality, where time and space
could slow down, speed up, bend, and even cease to exist, a reality
that transcends the limits of our daily perceptions, a reality that
makes no sense to us when we are thinking commonsensically.
Surely there are still new levels of reality that are hidden, and like
Einstein we ought to be curious to know what lies beyond our current
(commonsensical) understanding in physics.
As a student, I had mistakenly thought that physics was driven
mostly by mathematics and logical reasoning. Einstein’s conviction
was that principles are the driving force behind new discoveries,
while mathematics is necessary to make physics precise, to inform
the clarification of the principles, to explain and clarify our
characterizations of how we conceptualize phenomena, and to make
predictions. In short, math is not enough; it is a tool. The important
question is how does one come up with new principles? Einstein
answers: “Here there is no method capable of being learned and
systematically applied so that it leads to a new [principle]. The
scientist has to worm these principles out of nature by perceiving in
comprehensive complexes of empirical facts certain general features
which permit of precise formulation.”
He was saying that a scientist should make connections and see
patterns across a range of experimental outcomes, which may not
be related to each other in an obvious way. Once the scientist ekes
out these patterns, she makes a judgment call as to whether a new
principle of nature is necessary. But this is misleading. Facts are
statements about phenomena, but they don’t exist on their own; they
are always conceptualized, which means that they are, if only
implicitly, constructed theoretically. Experiments allow us to answer
theoretically constructed questions. Theory tells us what “facts” to
look for.
As an adolescent Einstein was free to play in his father’s electrical
company in Pavia, Italy. This play fertilized his imagination; it
enabled him to envisage what he would experience if he could catch
up to a light wave. His process of “worming” out these principles
entails visualizing phenomena that are not directly accessible to our
senses or current experiments. It eventually enabled him to
formulate theories that told us what we would find and helped us to
understand where we might look to find it.
How did Einstein know when to postulate his theory of relativity?
How, aside from his natural-born genius, was he able to arrive at his
principles? I found part of the answer in a lecture he gave at Oxford
University in 1933. “[The discovery of principles] are free inventions
of the human intellect, which cannot be justified either by the nature
of that intellect or in any other fashion a priori,” he said. But what
does Einstein mean by this? Sometimes to get around a scientific
problem, one must consider possibilities that defy the rules of the
game. If you don’t enable your mind to freely create sometimes
strange and uncomfortable new ideas, no matter how absurd they
seem, no matter how others view your arguments or punish you for
making them, you may miss the solution to the problem. Of course,
to do this successfully, it is important to have the necessary technical
tools to turn the strange idea into a determinate theory.
When I told the donors at National Geographic that I explored the
cosmos with my mind, I wasn’t jiving. Those theoretical physicists
who explored with their intellect, making “free inventions,” sounded
to me like masters of improvisation. Einstein gave me the hall pass
to continue my free inventions. But, like Einstein did, we must first
look to the fundamental principles underlying modern physics and
use them to explore some of the big mysteries physicists face. In the
pages that follow, we will engage in free inventions, trying to cook up
some new physics while journeying through some of the biggest
mysteries at the frontiers of cosmology and fundamental physics.
While some of the ideas presented in this book are debatable and
speculative, I hope that it nonetheless provides not only insight into
how a theoretical physicist dreams up new ideas and sharpens them
into a consistent framework but also, perhaps, the inspiration to think
of your own big ideas.
2

THE CHANGELESS CHANGE

A new idea comes suddenly and in a rather intuitive way.


That means it is not reached by conscious logical
conclusions. But, thinking it through afterward, you can
always discover the reasons which have led you
unconsciously to your guess and you will find a logical
way to justify it.
—ALBERT EINSTEIN

After many years spent developing my skills and ideas until they
were good enough for publication in physics journals, I finally
published my first independent paper in the Journal of High Energy
Physics. My article made an iconoclastic claim: Einstein’s cherished
idea of a constant speed of light could be violated in the early
universe if our actual universe were a three-dimensional membrane
orbiting a five-dimensional black hole. If this sounds like
gobbledygook to you, in hindsight, it is. But twenty years ago, such
subject matter was typical of what theorists worked on as they were
trying to integrate cosmology and string theory. I was especially
proud that the months of calculations I performed within the
framework of string theory provided these new solutions.
And so there I was, excited to give my first professional talk at a
picturesque university nestled in the mountains of Vancouver,
Canada. They were my calculations I was going to talk about, so I
knew them inside out, which contributed to my air of overconfidence.
It didn’t last long. Within five minutes of my talk’s beginning, I was
blasted with questions that soon transformed into a flood of criticism.
Attempts to continue my talk ricocheted against random comments,
delivered with a tone of unfriendliness, from the audience, attacking
the premise of the talk: “Why should we believe our universe is a
brane rotating around a 5D black hole?” I couldn’t help but feel
unwelcome and alienated. By the middle of the talk I stood dejected,
my fears of not being accepted as a peer erupting to the surface of
my mind. Just because your paper gets published doesn’t mean that
you will get into the club of physics. That day it felt obvious I hadn’t.
Then came a voice from the back of the room. The speaker was a
distinguished Indian physicist in his seventies decked out in a well-
groomed tweed suit. As soon as he began to speak, everyone shut
up, as if a demigod commanded his minions to silence.
The old man stood up and said, “Let him finish! No one ever died
from theorizing.”
It was the biggest lesson with the fewest words the audience and
I could have learned about the art of theoretical physics. Those
words would stay with me throughout my life as a theoretician. I took
the old man’s admonition as a reminder to never be afraid of even
the most absurd ideas, and to even embrace them. I finished my talk
without further interruption and even got a round of applause
afterward. Did I take my theory seriously years later? No, but the
exercise of journeying into a theoretical territory and then journeying
back has proven time and time again to be useful in surveying what’s
possible and, hopefully, what describes and predicts the real
universe. That moment was pivotal in my life and how I would
engage the art of theoretical physics for the next twenty years.
A year after that talk, after many failed attempts, I landed a job as
a postdoctoral researcher in theoretical physics at Imperial College
in London. The department had been founded by Abdus Salam,
who, along with Sheldon Glashow and Steven Weinberg, would win
the Nobel Prize in Physics for discovering a unified theory of the
weak nuclear interaction with electromagnetism. I was excited to be
following in Salam’s footsteps along the road to becoming a research
physicist. Yet somehow, despite my excitement, I quickly realized
that road was not what I thought it would be.
At Imperial, weeks and then months of work could pass with little
to show for it. If an idea did come to me, I invariably and quickly
discovered that someone had already developed and published it. If I
was performing a calculation, I would often hit a roadblock and have
to learn new mathematical techniques in order to tackle it. By the
time I learned the new math, someone would have already hit the
finish line and published the result before me. These experiences
forced me to wake up from my theory dreams to a reality in which
the prospect of becoming a scientist seemed dim. My contract was
for two years, but I relegated my expectations to another career,
perhaps going on the road as a jazz musician or teaching high
school physics, both admirable things to do. I would continue trying
my best, but ideas simply weren’t coming, and I would continue to
fake it and keep these frustrations to myself. I had everyone fooled.
Then one seemingly uneventful day, horror hit me. I received an
email from our theory group administrator that simply said:
“Professor Isham would like to speak with you.” I turned white like a
ghost. Chris Isham was the head of our theory group and I feared
that he had figured out that I was a fake. Everyone in our group
revered Isham for his exceptional abilities in quantum gravity and
mathematical physics. He was a tall Englishman with dark hair and
piercing eyes and who walked with a slight limp. Like his friend and
classmate Stephen Hawking, Isham suffered a rare neurological
disorder that kept him in constant pain. I had kept away from him in
fear of letting some gibberish slip out to ignite his physics bullshit
detector. Now I suspected and feared that he had figured it out on
his own, and my day had come to face him.
I decided to do a little preparation and read one of Isham’s
papers. Perhaps I could appease him. To my dismay, many of his
publications involved some of the most advanced concepts in math
and physics, with inscrutable names such as topos theory, quantum
logic, C-star algebras, and so on. I finally found a paper that he’d
written two decades ago that I could grasp. It was about the behavior
of quantum particles with half-integer spin, called fermions, in an
expanding gravitational cosmology. Electrons, quarks, neutrinos, and
most matter are examples of fermions, so it might seem a safe topic.
Still, I set off nervously for the meeting.
I tensely walked into his large office filled with books,
incomprehensible equations, and diagrams. On his desk an oddly
placed small statue of an angel faced a visitor. After a brief hello,
Isham didn’t waste time.
“Why are you here?”
I kept it real. “I want to be a good physicist.”
To my surprise, he said with a serious demeanor, “Then stop
reading those physics books!” Then he pointed to an isolated
bookshelf. “You see those books over there? They are the complete
works of Carl Jung. Do you know that Wolfgang Pauli and Jung
corresponded for decades? And Pauli’s dreams and analysis were
key to his discovery of the quantum exclusion principle.”
Isham revealed that he had been studying Jung over the last
fifteen years and had trained himself to do calculations in his
dreams. I couldn’t believe that I was hearing this from one of the
master mathematical physicists on the planet. Then he had a eureka
smile and said, “You know what? How about you come to my office
once a week? Write down your dreams and tell me about them.” He
suggested that I read Jung’s Volume I, book 9 entitled Aion:
Researches into the Phenomenology of Self as well as Atom and
Archetype, a collection of two decades of letters between Jung and
Pauli. At first, I was skeptical of the experiment. But I was also
feeling isolated in the theory group, and Isham’s invitation to talk
about my dreams was an opportunity to spend quality time with one
of my physics idols.
Our weekly discussions started with me telling him about random
dreams that had no apparent relation to physics, such as those
about past relationships that continued to taunt me. During our time
together, Isham would share his perspectives on some of the
mysteries that our field faced. One of those was the problem of time
in quantum gravity. While our physical (and psychological)
experience of the flow of time is taken as fact, time disappears in the
equations of quantum gravity. Isham worked on this problem and
was a proponent of a new notion of time called internal time. It was
no surprise to me to learn that these ideas were inspired by his
exploration of psychology and mysticism.
As the weeks passed, I told Isham about what I thought was a
trivial dream. In Jungian philosophy, dreams sometimes allow us to
confront our shadows with the appearances of symbols called
archetypes. I saw one here. I was suspended in outer space and an
old, bearded man in a white robe—it wasn’t God—was silently and
rapidly scribbling incomprehensible equations on a whiteboard. I
admitted to the old man that I was too dumb to know what he was
trying to show me. Then the board disappeared, and the old man
made a spiraling motion with his right hand. Isham was captivated by
this dream and asked, “What direction was he rotating his hands?” I
was baffled as to why he was interested in this detail. But two years
later, while I was a new postdoc at Stanford, I was working on one of
the big mysteries in cosmology—the origin of matter in the universe
—when the dream reappeared and provided the key insight to
constructing a new mechanism based on the phenomenon of cosmic
inflation, the rapid expansion of space in the early universe. The
direction of rotation of the old man’s hand gave me the idea that the
expansion of space during inflation would be related to a symmetry
that resembled a corkscrew motion that elementary particles have
called helicity. The resulting publication was key to earning me
tenure and a national award from the American Physics Society.
Chris Isham’s method proved to work for me. But he and I weren’t
alone here. It turns out that some of the biggest breakthroughs in
science were inspired by dreams, including Einstein’s theory of
relativity.

Beginning when Einstein was a teenager hanging out in his father’s


electric lighting company, he would play with imaginations about the
nature of light. He would try to become one with a beam of light and
wondered what he would see if he could catch up to a light wave.
This matter found itself in the playground of Einstein’s subconscious
and revealed a paradox in a dream. It is said that Einstein dreamt of
himself overlooking a peaceful green meadow with cows grazing
next to a straight fence. At the end of the fence was a sadistic farmer
who occasionally pulled a switch that sent an electrical current down
the fence. From Einstein’s birds-eye view he saw all the electrocuted
cows simultaneously jump up. When Einstein confronted the devious
farmer, there was a disagreement as to what happened. The farmer
persisted that he saw the cows cascade in a wavelike motion.
Einstein disagreed. Both went back and forth with no resolution.
Einstein woke up from this dream with a paradox.
In the account of Einstein’s dream, and other accounts of the role
of dreams in creative work, such as music, science, and visual art,
there is a common theme: a paradox is revealed through
imaginations that are contradictory in the awake state. It’s as if the
mind’s eye can access an intuition beyond the waking state and not
restrict our imaginations to the self-editing that our conditioning might
impose during the waking state (unless you’re a great daydreamer).
Perhaps dreams are an arena that can enable supracognitive
powers to perform calculations and perceptions of reality that may be
incomprehensible in our wake state. In my case, my paradox was
making an equivalence between incomprehensible equations
presented by the bearded man and his counterclockwise whirling
hands. This counterclockwise motion turned out to summarize the
mathematics that was obscuring the underlying physics to be
unveiled.
My preoccupation with equations as the way to access deeper
physical reality was confronted by this paradox. I discovered years
later that I was not alone. My friend, virtual-reality pioneer, composer,
multi-instrumentalist, and author Jaron Lanier told me that in his pre–
virtual reality days, he used to hang out with Richard Feynman at
Caltech. During that time Feynman was experimenting with other
ways of doing physics, including using his body to intuit physics. This
got the young Lanier thinking about how the human body could
interact with computers in new ways and inspired what would
become VR. Einstein’s paradox asks how it could be possible for
both realities of light to be true. He ultimately resolved that paradox,
and in doing so would uncover one of his most fundamental
principles, one that was key in discovering not only his theories of
special and general relativity but also the nature of all known four
forces!
These thought experiments with light also led Einstein to other
paradoxes, such as in James Clerk Maxwell’s theory of
electromagnetism, which mathematically describes the motion of
light waves. In this case, Einstein found inconsistencies that forbade
absolute rest for a moving wave of light (an electromagnetic wave).
In his groundbreaking 1905 paper On the Electrodynamics of Moving
Bodies, Einstein resolves this inconsistency by elevating the role of
principles in physical law. During Einstein’s time, physicists assumed
that the universe is filled with a substance called the luminiferous
ether that light could move through, like water waves moving through
the medium of water. Mainly because of his intuition that an
electromagnetic wave of light could never come to a complete stop,
Einstein gave up the ether. Based on years wrestling with a handful
of conundrums with electromagnetic theory, Einstein made an
intuitive leap and postulated the principle of the invariance of light:
The Principle of Invariance of Light: The speed of light in
empty space is the same for all inertial observers regardless of how
fast they are moving relative to each other.
It’s worth saying a little about some of the physical reasoning that
justified Einstein’s adoption of this principle. In his groundbreaking
paper The Electrodynamics of Moving Bodies, Einstein finds a
paradox in Maxwell’s equations that describe electric and magnetic
fields. Both electric and magnetic fields exert forces on charged
particles. It was well known since Galileo’s time that mechanical
forces would be unchanged (invariant) for observers that are moving
at constant speeds relative to each other. However, Einstein found
that this was not the case for Maxwell’s equations. For example,
different frames of reference would give different physical results for
electric and magnetic forces. Einstein realized that this could be
fixed if the electric and magnetic fields would also change into one
another depending on the frame of reference.
Visible light is a manifestation of electromagnetic waves at a
given set of frequencies that our eyes can detect. Einstein realized
that if electric and magnetic waves traveled at a constant speed, the
electric and magnetic force laws would remain intact. Einstein found
that the transformations that related these moving observers, which
are known as Lorentz transformations after the mathematician who
first found them, did something surprising: they warped spatial and
temporal measurements of relative observers. The consequence of
Einstein’s insight is that we know electromagnetic waves are unified
in four dimensions, three spatial and one temporal. And this relativity
extended to all known laws.
Einstein realized that if time depended on the observer’s velocity
it would be possible to retain the invariance of light in different
moving frames. A similar argument was made for the length of
objects in relative frames of reference. Therefore, both time and the
sizes of objects are no longer absolute but are subject to change
depending on how fast one is moving relative to another inertial
frame of reference. The lesson here is that the invariance of the
speed of light required time and space to be relative—we need
relativity to have absolutes.
In our everyday experience, we simply do not experience time
slowing down for us when we move faster. This is because the
velocities we are used to are not large enough to see the relativity of
time and space. Nevertheless, Einstein’s new theory made these
bold predictions, and it was up to experimentalists to find a way to
put them to the test. Not far from where I lived in Hanover, New
Hampshire, when I was a member of Dartmouth’s physics
department, the phenomenon of time dilation was finally confirmed
by David Frisch and James Smith, who attacked the problem with
the help of a particle called the muon. The muon is the cousin of the
electron. Like the electron, a muon has a negative charge that we
describe as −1, but there is a critical difference: a muon has two
hundred times the electron’s mass. The muon is also unstable and
will decay into an electron at a fixed rate of two-millionth of a second.
Therefore, we can treat the decay of a muon as a standard clock. In
1941 Bruno Rossi and David Hall made use of the serendipitous fact
that muons travel to Earth from outer space close to the speed of
light to get a rough estimate of how moving at such a high speed
dilates time. They measured the half-life of these rapid muons at
high altitude and found, as Einstein had predicted, that these muons
lived longer than muons that were at rest. In 1963 Frisch and Smith
made much more rigorous measurements of the speed of
extraterrestrial muons, which enabled an even more precise
confirmation of how high speed warps time. These observations
were refined and continued to be confirmed with the precise
numerical predictions of Einstein’s theory of special relativity. But
there is something even more special about special relativity.
Invariance isn’t just an empirical fact. The core idea behind
invariance is based on the symmetry properties of geometric objects.
The key insight in Einstein’s invariance principle is that physical
entities like space-time and the other forces are in direct
correspondence with geometry. It still baffles mathematical physicists
why geometrical structures are linked with space-time and the
building blocks of matter; no one has expressed this better than
Nobel laureate Eugene Wigner, who wrote that “The miracle of the
appropriateness of the language of mathematics for the formulation
of the laws of physics is a wonderful gift which we neither
understand nor deserve.”
Consider the transformations of a square that leave it invariant.
These are its symmetries. Rotate the square by ninety degrees and
it looks the same. Rotate it again by ninety degrees and it still looks
the same. Ultimately there are four successions of ninety-degree
rotations that would leave the square looking the same. We can also
flip the square about its horizontal center, and it looks the same. We
can flip it about its vertical center and it also looks the same. We can
flip the square across a diagonal line in two different ways and it
looks the same. Therefore, there are eight different ways we can
change the square and it looks the same. Said another way, there
are eight transformations that leave the square invariant. The square
is geometrically described as a two-dimensional object that has an
eightfold symmetry. If I wanted to communicate with aliens that could
not see, but knew numbers, I could describe many geometric objects
by their symmetry transformations.
So, invariance captures an idea that science writer K. C. Cole
describes as the differences that don’t make a difference. In the case
of special relativity, I can move as fast or slow as I want, and the
speed of light will not change. And this means that we can identify an
underlying geometric symmetry. What would a world like that look
like geometrically?

FIGURE 1: The eightfold symmetry transformations that leave the square


invariant.

In special relativity Einstein found that there is a symmetry in all


inertial coordinate systems that leaves the speed of light the same.
The key to being able to enforce that the speed of light is the same
in different inertial coordinate systems is to find the correct object, or
observable, to relate coordinate systems to each other. The
observables that we often use are distances and time. In three-
dimensional space, distances alone are not sufficient because the
position of a moving object can also have a direction. The
mathematical object that describes distances with a direction is
called a vector. Einstein realized that he could unite space and time
into a four-dimensional vector, a space-time vector. This vector has a
length and direction in both space and time. But to do this, we must
put the time dimension on the same footing as a spatial dimension.
The hack: multiply time by a velocity and we get a dimension of
length.1 In this way the time dimension resembles a spatial
dimension. Being equipped with a four-dimensional space-time
vector enables us to relate them to each other in different coordinate
systems. These relations or space-time vectors between different
coordinates are exactly analogous to rotation transformations of
points on a circle.
Consider an idealized Ferris wheel ride, where one person, Kolka,
is on the Ferris wheel and another, Jim, is on the ground. Before the
ride starts, both Kolka and Jim seem to share the same coordinate
system, which they can label as points on an x-y plane. But as the
Ferris wheel and Kolka start rotating, her coordinate system will be
labeled by a rotating coordinate system. But even though both riders
will disagree about their respective coordinates, they will agree on
the invariant, which is the length of the radius of the Ferris wheel. In
the same spirit, Einstein’s theory of invariance relates coordinates
that differ from each other by relative velocities but retain their four-
dimensional “radius” that remains unchanged by all observers. The
compromise for preserving the four-dimensional radius is that
observers that move at velocities comparable to the speed of light
will experience the rate of time slowing down and lengths shrinking
relative to observers that are at rest.
The four-dimensional space equipped for four-dimensional
vectors was invented by Hermann Minkowski, who was inspired by
his former student Einstein and realized that the system of relativity
had an underlying geometry that is a four-dimensional space-time
continuum that had a symmetry that left the speed of light as an
invariant. The constant speed of light is formulated as the geometry
of a cone embedded in a four-dimensional space. In this geometry a
ray of light traces out the cone, and the slope of the cone represents
the constant velocity. And just like the radius of a sphere remains the
same length regardless of how one moves on the surface of the
sphere, in this four-dimensional space-time the light cone would
remain the same no matter what frame of reference is used.

FIGURE 2: The light-cone structure of Minkowski space-time.

The triumph of special relativity was based on a simple and deep


new principle of invariance, but the principle was so deep and
universal that importance of the idea of invariance did not stop with
special relativity. As we will soon see, the invariance principle
became the master organizing principle behind all the physical
forces, from all nuclear forces to gravity. Einstein further
hypothesized that the laws of physics should not care about any
relative state of motion, including those in which observers are
accelerating relative to each other. If you are spinning in a rocket
ship out in space or sitting still on a planet, the laws of physics will
function identically. This led Einstein to perceive a reality beyond our
five senses to discover a new law of gravity, the general theory of
relativity, by deepening his invariance principle to observers
undergoing any type of motion. He came to this realization by what
he said was “the happiest thought of my life.”

General relativity is in hindsight a logical, but radical, step from its


cousin, special relativity, which assumed the invariance of the laws
of physics between observers that moved at constant speeds
relative to each other: the jargon for this is that they are inertial
observers. Those observers were all in reference to a fixed space-
time structure, Minkowski space. But Einstein’s insight was to make
the laws of gravity invariant for observers situated in wildly different
space-time structures. This insisted that space-time can fold, warp,
stretch, or generally be dynamical like an electric field. We are now
ready for the master principle.
The General Principle of Invariance: All systems of reference
are equivalent with respect to formulation of the fundamental laws of
physics.
Einstein’s principle of invariance proved so powerful in the
discovery and underpinning of special relativity that he used it to
unearth the hidden reality of space-time and the gravitational force,
which we know as the general theory of relativity, but it would also
prove to be more influential again. As we will see, this turn of the
tides, of placing symmetries as a central principle, would also
become the driving force to writing down the other nuclear forces.
Einstein arrived at his theory of general relativity by a thought
experiment—his happiest thought—by realizing that the experience
of an accelerated person in the absence of a gravitational field (like a
person inside a rocket ship accelerating in outer space) is the same
as a person standing at rest on a planet with a gravitational field. The
converse is also true. If you are in free fall in the presence of Earth’s
gravitational field, your acceleration can “erase” the earth’s
gravitational field. In other words, if a person is isolated in an
elevator and is in free fall toward Earth, they would feel no
gravitational forces.2 Einstein reasoned if acceleration could mimic
the effect of a gravitational field, then gravity itself has to account for
both acceleration and a gravitational field on the same footing. In
other words, the law of gravity has to respect a symmetry between a
constant gravitational field and accelerated motion without a
gravitational field. How does one formulate such a symmetry? The
strategy is to use what worked before, that is, special relativity.
General relativity requires invariance of the laws of physics
regardless of the relative motion of observers. This requirement can
be restated into a more general symmetry relating any type of
coordinate transformation, called general covariance. For instance, a
rotating observer will experience the laws of physics the same as a
nonrotating observer.3 The compromise for putting all observers on
the same footing gives two radical features of gravitation. The first is
that the gravitational force is not a force in the way we’re used to
thinking about it, but a feature of the curved space-time. In the
Newtonian world an electric field could change the direction of a
charged particle. In general relativity, particles will move in the
straightest path possible. But in a curved surface, such as the
surface of a sphere, the shortest distance between two points is not
a straight line.
Imagine that you lived exactly at the equator of the earth and
wanted to travel around the world using the shortest path to return
home. The shortest path would be exactly a horizontal curve that
circumscribes a great circle around the earth. This shortest path is
not a straight line. Likewise, if space is curved then the shortest
distance won’t necessarily be a straight line in flat space. These
straightest paths are called geodesics, and what we call the
gravitational force is simply a geodesic motion in a curved space that
a star, for example, bends.4 The motion of Earth moving around the
sun is nothing more than a geodesic due to the warping of space-
time by the sun, and our planet is just moving around the contours of
the curved space-time.

FIGURE 3: The intrinsic curvature on the surface of the sphere is


measured by dragging and keeping parallel a flag around a closed circuit.
The space is curved if the dragged flag fails to be aligned with its original
orientation.

To do physics we need dynamical equations for the objects of


relevance. For electromagnetism, the electric and magnetic fields
obey a set of four equations codified by James Clerk Maxwell. The
principle of invariance allows us to repackage all four equations into
one equation that is invariant under the coordinate transformations of
special relativity.5
I suffer from having a bad memory and always forget one of
Maxwell’s four equations. So I was thrilled to realize that I can use
the principle of invariance to write down one equation that contains
all the other Maxwell equations.6 And using the same approach,
Einstein figured out that there is an object that can codify the
dynamics of space-time that respects the general coordinate
transformations. Serendipitously this mathematical object was
already discovered by mathematician Bernhard Riemann. The
object, called the Riemann curvature tensor after its inventor,
captures the dynamics of curved space-time in the presence of
matter and energy. The famous Einstein equations of general
relativity describe how the curvature of space-time is warped by
matter and energy. And similar to the one Maxwell equation that
contains others, the singular Einstein equation unpackages ten
dynamical equations for the field of space-time.
Perhaps the most profound consequence of Einstein’s principle of
invariance is that space-time itself is not some empty, fixed stage but
rather is a dynamical field. But this leaves an interesting question.
Since the gravitational field is space-time, when the gravitational field
vanishes, does space-time itself cease to exist? In other words, what
is no-space? We will revisit this question, but answering it will require
that we meet our other principles first. Let’s jump into the quantum.

FIGURE 4: Three black holes affect the dynamics of space-time by


warping the geometry of the space-time. This malleability of space-time
has a range of physical effects including but not limited to gravitational
forces and bending of light.
3

SUPERPOSITION

If you think you understand quantum mechanics,


you don’t understand quantum mechanics.
—RICHARD FEYNMAN

It was sophomore year and the moment we physics majors had


been waiting for. The last year and a half had been a tedious grind of
problem sets that involved boring blocks sliding down inclined planes
and systems of masses hanging on pulleys. But that day we sat
attentively. It was our first quantum mechanics class, and we were
waiting for Professor Lyle Roelofs, a tall man with a thick mustache
in a plaid shirt, to deliver his first lecture. He began by saying, with a
straight expression on his face, “I can assure you that after learning
quantum mechanics, there will be no need to drop acid.” I thought he
was joking. Twenty-eight years later, however, I can report that he
was right. Just when you think you understand it, when you stop to
think of what quantum mechanics is saying about reality, there is
really no need to drop acid—your mind has already experienced
alternate reality.
Quantum physics underlies our electronic technology, and
although we think about it as the physics of the very small, it is likely
to apply to galaxies, too, as it probably describes the behavior of
whatever particles make up dark matter. And, as we’ll see, quantum
mechanics is probably central to the existence of the universe, too.
In fact, one of the reasons I became a physicist was Stephen
Hawking and James Hartle’s proposal of the wave function of the
universe, which posits that the entire universe is a quantum
mechanical system.
Perhaps it comes as no surprise that, just as no one ever
described the general theory of relativity as being like an acid trip,
quantum mechanics wouldn’t offer up an obvious central principle
like the principle of invariance. That doesn’t mean we can’t try to find
one. If I had to choose one (which I guess I do!), it would be the
principle of superposition of states.
Of all the strange things about quantum mechanics, I believe
superposition is the main point of departure of the microscopic
quantum world from the ordinary world of large things (like dogs and
airplanes) of classical physics that we experience. Once we have
this principle, it invokes strange features such as nonlocality, the
uncertainty principle, and complementarity, which, if you’re familiar
with quantum mechanics, are all pretty strange, and if you’re not,
you’re about to! These will form the basis of much of the cutting-
edge mysteries we are currently wrestling with and exploring in these
pages.
To appreciate the bizarre world that the principle of superposition
of states describes, we should first understand the concepts of
superposition and state independently. In classical mechanical
systems, the physics of the macroworld you inhabit, every object has
two attributes that describe what physicists call its physical state.
One is its position, which is where the object is. The other is the
object’s momentum, which is defined by its mass and its velocity.
This means that momentum is a vector, and includes both the speed
of the object and the direction it is traveling in, as well as the object’s
mass. We can think about the physical state as a point in a two-
dimensional space, where the y-axis denotes the momentum and the
x-axis denotes the position. This space is called phase space. Phase
spaces are key tools in analyzing all dynamical classical systems,
whether a baseball or a spaceship—or even classical particles.1
Each occupies a unique point in phase space. So if we have
something simple, such as a ball in motion, we are able to predict
with certainty what its state in the phase space will be in the future,
once we apply to it the dynamical laws in classical physics. This is
called determinism. A classical object can only be at a unique
position and velocity at a given moment of time. I hope so far this
makes good common sense.

FIGURE 5: A phase-space diagram showing a particle’s momentum and


position.

You might think that, if states make good common sense, then it
must be superposition that doesn’t. But it can make common sense,
too. In fact it is a common and useful property of waves. If we drop a
stone on a calm pond, a periodic wave will propagate radially
outward from where the stone hit the water. If we drop other stones
nearby, other periodic waves will also propagate radially outward
from those points as well, and sooner or later those waves will
encounter each other and combine. This will result in a more
complicated pattern of waves, where they can either add up or
subtract from one another. The ability of waves to combine to make
more interesting waves is at the heart of superposition: in fact, the
word superposition literally means the process by which simple
periodic waves of differing frequencies are added up to make a
complicated-looking wave. An intricate-looking water wave is made
up of billions of water molecules that collectively superpose to take
the shape it has. This isn’t just something that happens in bodies of
water; it’s why there can be dead spots if stereo speakers aren’t set
up correctly, and it is also how electronic music synthesizers work.
Conversely, a complicated-looking wave can be decomposed—
mathematically, anyway, if not in a pond—into a large collection of
periodic simpler waves.

FIGURE 6: To the left, solid and dashed periodic waves can superpose
(be added) to form a more complex wave to the right. Conversely, the
right-hand wave can be decomposed into the two periodic waves on the
left.

The ability of waves of water atoms to superpose made it all the


more confusing for physicists when they first encountered the bizarre
behavior of electrons in their experiments. In a nutshell, individual
electrons were thought to be particles, but they exhibited wavelike
behavior instead. Quantum mechanics was invented, in part, to deal
with those experiments.
Here is the punch line: A single electron can exist as a
superposition of many states at the same time. In other words, a
single electron can be in many places, or have many momenta, at
once. More generally, behavior of electrons or any quantum entity
can be stated as follows:
The Quantum Superposition Principle: A quantum system is
expressed by adding many distinct states at the same time.
A common example of quantum superposition concerns the
position of an electron moving in space. The superposition principle
says that the state of an electron’s unique position in space is
equivalent to the same electron having a wide array of velocities at
the same time. The converse is also true. If the electron has a state
describing it to have a given velocity, it can equally be expressed as
a superposition of one electron having a very large range of
positions.
You can see why Professor Roelofs thought this was crazier than
tripping on acid: all your physical intuition about the classical world
should go out the window. How can one electron be in several
places at once? How can one particle be moving at many different
velocities at the same time?
One of the most mysterious experimental displays of the quantum
strangeness of superposition is the double slit experiment. Richard
Feynman calls the double slit experiment “impossible, absolutely
impossible to explain in any classical way, and has in it the heart of
quantum mechanics.” Here is how it goes: One electron at a time is
shot toward a screen, which lights up where the electron lands.
Between the electron gun and the screen is a barrier with two small
holes that allow the electrons to pass through. At first, we see
nothing out of the ordinary, just electrons landing and lighting up the
screen like fireflies. After a large number of individual electrons
arrive on the screen, however, they distribute themselves in a bizarre
way. Since the electrons are particles, we expect them to pile up in
two bundles directly behind the holes. However, the electrons land
on the screen to create what’s called a wave interference pattern,
with areas with lots of light spots and areas with few. How did the
individual electron particles know to avoid the location of the
previous electrons to collectively exhibit a wave pattern later? What
exactly is the electron doing when it arrives at the double slit? Is it
behaving like water waves that go through both holes and interfere
after they pass through?
We could partially explain the interference pattern if we assume
that the electrons behave like a wave as it approaches the two
nearby holes. The electron wave would split up and pass through
both holes and recombine forming an interference pattern.
Experimenters tried to find out if that’s what was going on by putting
detectors at one opening to see where the electron went through.
What they saw instead was that the interference pattern on the
screen completely disappeared. So, when we are not looking at the
electrons during their flight, they behave like waves, and when we try
to pin their positions down, they lose their wavelike properties and
collapse into one location out of the many possible places they could
have been. It’s as if the electrons knew when they were being
watched to change their wavelike properties. Ever since this bizarre
situation presented itself, many physicists have attempted to
reconcile it with either modifying quantum mechanics or giving it a
new interpretation. There is still no general agreement!
Although experiments like the double slit experiment were and
still are quite shocking, it turns out that most of our modern
technology stems from the quantum mechanical understanding and
ability to control the electron. My favorite is the solid state laser.
Another is the nuclear magnetic resonance technology, which drives
the MRI machine used to do brain scans in hospitals. Electrons carry
a negative electric charge and have another property that physicists
call half-integer spin. Spin is called that because, mathematically, it’s
something like the spinning of a top. Tops spin clockwise and
counterclockwise; electrons have elementary spins we call up and
down, and they behave like tiny magnetic north and south poles.
Spins can be flipped by the application of a precise magnetic field,
and this is at the heart of an MRI.2
The double slit experiment also inspired those in other disciplines,
including both neuroscience and mysticism, with the behavior of the
observed electron indicating that perhaps there is some
undiscovered link between quantum mechanics and consciousness.
These speculations continue to stir mystery, controversy, and
confusion, even among Nobel Prize–winning physicists, and there
may even be some truth to these intuitions. I will present my own
suppositions later, but for now, it is important to see that this weird
behavior stems from two ideas: that the electron’s location in space
is a superposition of position states, and when we make an
observation we can only see the electron realize itself at one
position.
This property is seen in a wide array of quantum mechanics and
is currently being exploited to make quantum computers a reality.
This conundrum is at the root of an issue in quantum mechanics
called the measurement problem, a subject we must confront. The
bizarre quantum superposition could also be a useful attribute for a
superhero. Like many youngsters, I was hooked on Marvel comic
books and imagined creating my own superhero. It turned out that it
didn’t matter that much that Stan Lee and I went to the same high
school. Also, my drawing skills weren’t that great, so I didn’t end up
pursuing that path. I am now amazed to see how quantum physics
has entered in Marvel movies, such as Ant-Man and The Avengers.3
If I were to create a new character, I can think of a few properties of
the quantum world that would fit the bill of major new attributes for a
superhero. If my hero was fighting many villains at once, then he or
she would have a major advantage over a nonquantum hero. A
quantum superhero with the ability to be superposed in many
positions at the same time could fight many villains at once. Of
course, this raises some deep questions. If an electron causes
trouble for us once we interact with it going through a slit, what’s
going to happen when the quantum hero interacts with classical
villains? We will return to this issue when we discuss the
measurement problem inherent in quantum mechanics.
But first it’s useful to explore more quantum weirdness.

In 1923 a French prince named Louis de Broglie claimed that all


matter possesses wavelike properties. First, by paradoxical fiat, de
Broglie related the momentum of an electron to its wavelength, and
further that all matter—anything with mass and momentum—also
has a wavelength associated with it. This hypothesis was used to
correctly explain one of the great mysteries of early twentieth-century
physics, which is why negatively charged electrons don’t go crashing
into the positively charged nucleus of atoms. If we identify the wave
as the electron’s orbit around the nucleus, then the longest
wavelength predicted by its mass would correspond to the closest
orbit that it could make around the nucleus. The electron would not
be allowed to go to lower orbits and fall into the nucleus because no
lower wavelength orbitals would exist for the wave. This picture was
also consistent with Niels Bohr’s model of the atom, and why he had
hypothesized that electrons existed in orbits that could only have
discrete, integer values. This model correctly predicted and
explained the absorption and emission spectrum of gasses that
classical physics could not explain. Despite having a host of models,
such as the Bohr atom and de Broglie’s orbits, to account for the
wavelike behavior of electrons, physicists still lacked a precise
equation to describe what was going on. At least they did until Erwin
Schrödinger returned from a short ski trip in the Swiss Alps with a
most beautiful and elegant dynamical wave equation describing the
wavelike superposition of quantum matter.
Schrödinger is one of my heroes in physics, and he inspired many
young physicists of my generation as well. He wasn’t the
stereotypical “geeky” physicist that is often portrayed in movies and
in TV shows like The Big Bang Theory. He had a host of other
interests such as poetry, color theory, and Eastern mysticism, and I
am convinced that his polymath tendencies influenced his creativity.
While no one knows exactly how he came up with it, Schrödinger
discovered a differential equation—which we call the Schrödinger
equation—for all quantum systems. It is an intriguing fact that all our
theories are described by differential equations; for some reason,
after all the elegant formulations of a physical theory, when the
rubber hits the road, physicists end up solving differential equations.
The solution of the Schrödinger differential equation is a
mathematical function that contains all information, including
superposition of states, about quantum systems: the wave function
ψ.4
The Schrödinger equation is so aesthetically beautiful, I can’t help
but write it down.

This singular equation determines the behavior of the wave


function, denoted by the variable, spoken as psi (and sounds like
sigh). It describes how the wave function of a quantum system
interacts, and changes in time. This equation predicts all quantum
phenomena, including the periodic table, physics of a neutron star,
and even the semiconductor device in your smartphone. The
equation simply says that the temporal change, or evolution, of a
quantum wave function (right side) depends on the energy, which
encodes interactions acting on it (left side). Once we know the
evolution of the wave function, we get new insights into and
predictions for the quantum system at hand. Some of these
properties led to the invention of gadgets like transistors, the heart of
all computer-driven technology. The superposition principle emerges
naturally from it. This is because the Schrödinger equation is what is
known as linear. This means that if one finds two independent
solutions of the wave function, then the sum of those solutions would
also be a solution. Periodic waves obey this mathematical property
of linearity. Thus, the wave function describes waves that can be
decomposed into an infinite set of simple harmonic waves of differing
frequencies.
So what is the electron doing, according to the Schrödinger
equation? Schrödinger interpreted the electron as a highly peaked
material wave, like a pulse: the wave is the electron. But with this
interpretation a natural paradox presents itself: If a highly peaked
wave function is to describe a single electron, what does it mean
when the electron wave function hits a barrier and splits in two? One
common feature of a moving wave is that if it starts off highly
peaked, at some later time it will spread. But what becomes of the
particle when the wave loses its shape and splits into many pieces?
Are we to think that the electron will split up into many electrons?
Precision experiments from particle accelerators reveal that the
electron is an elementary particle and is not a composite of any
smaller unit of matter. So if the wave splits, the electron certainly
does not split in two. It could be that the electron does lose its
identity when it encounters an environment that enables it to shape-
shift. We will return to this issue. For now, let’s accept the electron as
an indivisible and independent unit.
Schrödinger’s interpretation that the wave function represents a
material wave whose shape describes the position of a particle was
problematic, and it forced physicists at the time to seek an
interpretation of the wave function to also make sense out of a
quantum particle being in many states at once, as is suggested by
the double slit experiment. Max Born, another leading quantum
physicist, provided an ingenious interpretation of the Schrödinger
equation that dealt with the problem of a splitting wave: he argued
that the wave function is not a material wave but a wave of
possibilities, so that where the wave has the largest amplitude the
particle is more likely to be there, and correspondingly less likely to
be where the wave’s amplitude is lower. Applying the Born
interpretation to the double slit experiment, the wave pattern we see
is the result of a probability wave that goes through both slits. When
we look to see which hole the electron is at, the probability
superposition collapses into one position outcome and the
interference pattern disappears. This still doesn’t solve every
problem with the double slit experiment. You should scratch your
head and wonder how the act of observation collapses this
probability distribution in the wave function to the observed value.
Also, what does it mean for a probability wave to go through the slit?

The competing interpretations of the wave function forced the


architects of quantum mechanics to take philosophical stances about
how to interpret quantum mechanics generally. Albert Einstein, one
of the founders of the theory, took a realist stance. Realism demands
that there is an objective physical world that is independent of our
existence. This seems reasonable since the universe had existed
billions of years before stars, planets, and humans came on the
scene. Saying that a probability wave goes through a hole avoids
saying exactly what the electron is actually doing in that region of
space. In accounting for the double slit experiment, a realist posits
that the quantum theory should explain what the electron does when
it encounters the two open slits and the resulting interference
pattern, with or without the presence of a measuring interaction.
There is a realist interpretation that does this, the de Broglie–
Bohm pilot wave. In the de Broglie–Bohm theory there is no
superposition in the wave function, but a particle and an invisible
wave that propagates through the double slit. This electron surfs the
contours of the wave and ends up tracing out the interference
pattern of the wave. This de Broglie–Bohm interpretation still uses
the Schrödinger equation but trades off the probability weirdness for
another kind of weirdness, which allows for nonlocal interactions
between physical objects. In the de Broglie–Bohm theory, the wave
can cause faster-than-light correlations between different electrons.
If the wave changes in one region, an electron motion will be
instantaneously affected at a distant region.
FIGURE 7: The double slit experiment as represented by the de Broglie–
Bohm pilot wave theory. The lines that emanate from the two slits are the
trajectories of the electrons along the contour of the nonlocal quantum
potential surface.

Niels Bohr and his followers rejected realism and actually


formulated quantum mechanics, the form that is taught in most
textbooks, to avoid asking questions about the electron’s reality
when it’s not being observed. Some such as Werner Heisenberg and
Born even questioned the existence of particles until an interaction
takes place. Influenced by Eastern philosophy such as Buddhism
and Taoism, Bohr believed that our classical experience and
perceptions were incompatible with the intrinsic duality that quantum
entities possessed. According to Bohr, quantum entities possess
pairs of contradictory qualities, such as wave and particle, energy
and time, electric and magnetic, spin up and spin down. In Taoism
the yin and yang represent contradictory or opposing qualities that
coexist to describe the whole. Likewise, the whole electron
possesses both the contradictory wave and particle properties at the
same time. It is through our macroscopic measuring devices that we
see either the particle or the wave properties revealed, and not both.
Bohr elevated this duality to a principle he called complementarity.

FIGURE 8: A visual representation of complementarity. There appears to


be two independent images. Upon closer inspection, the boundaries of
both images define each other.

Some of Bohr’s brilliant young followers incorporated his


complementarity principle into the actual mathematical formalism of
quantum theory. The Nobel Prize was awarded to a young Werner
Heisenberg, who formulated the uncertainty principle, which is
rooted in complementarity. To see this a bit more clearly let’s
combine Bohr’s wave-particle complementarity with the
superposition principle to see how uncertainty follows naturally.
According to the superposition principle, simple periodic waves of
different frequencies can be added up (superposed) to give a more
complicated wave. I can add up many periodic waves of differing
frequencies to give a pulse, which approximates a particle like the
electron. A quantum state described by a wave with definite
frequency is actually the same as a particle with a definite velocity.
This happens because the speed of a periodic wave is proportional
to how rapidly it oscillates—its frequency. And according to the
superposition principle it takes a very large superposition of many
different velocity states to approximate a state of definite position of
a particle.
This feature of duality and uncertainty marks a profound
difference between the classical and quantum worlds. In classical
systems, the dynamics are given in pairs of physical quantities, for
example, as we discussed position and momentum. These dual
pairs come together to form phase space, and in principle we can
know everything about the position and momentum at the same time
with complete accuracy. On the other hand, in quantum mechanics
we can only know one of the quantities with precision, at the
expense of randomizing the other. The key idea is that dual qualities
like position and momentum simply and inevitably disrupt each
other’s certainty.
Heisenberg came up with a thought experiment, the Heisenberg
microscope, to highlight this compromise. He imagined that if we
observe an electron, light has to bounce off it, exchanging a definite
amount of momentum, which will change the electron’s trajectory,
hence randomizing its subsequent position. Consider a state with a
unique definite frequency. I ask you, where is the wave? If I pick a
point on the wave, it looks exactly the same as any other point if I
shift along by one complete cycle, so any point on the wave is not a
unique position. Therefore, the specific location of the wave is
indeterminate. A perfectly periodic wave is everywhere! Now imagine
a pulse, similar to what we see on the devices that measure your
heartbeat. The position of the pulse is located where the pulse is
highest in height. A pulse is a wave that is very localized in space.
However, the frequency of a pulse is indeterminate because a
frequency is defined to be a quantity that depicts how often a wave
repeats itself in a given amount of time. The singular pulse only
repeats itself once. So, a periodic wave has a definite frequency but
an indefinite position. And a pulse (now think of it as a particle) has a
well-defined position but an indefinite frequency. There is a tradeoff.
Quantum states that are localized in space are indefinite in their
momentum, and those with precise momentum have delocalized in
space.
Einstein was vehemently against the idea that quantum
mechanics, and so nature, was fundamentally operating on chance.
He famously said, “God does not play dice,” reflecting his realist
stance on physics. Bohr responded, “Don’t tell God what to do.”
Despite the witticisms and counterarguments, Einstein persisted in
disagreeing with Bohr and his followers and set out to find the death
blow to the anti-realism of quantum mechanics. Einstein teamed up
with his own duo of young researchers, Boris Podolsky and Nathan
Rosen. Together, they wrote a paper containing what is known as the
EPR paradox. EPR showed that special states exist in quantum
mechanics such that two particles can have opposing features; these
are called entangled states. A version of this thought experiment
involved a pair of photons that is created by the decay of a particle at
rest. If the photons start out in a spin-zero state and fly off in
opposite directions at astronomical distance, the measurement of the
spin of one photon will predict with 100 percent probability that the
other spin has to be opposite, because the photons are in a spin-
zero entangled state. The other spin will be communicated
instantaneously. EPR coined this as “spooky action at a distance,”
and it was seen to be in opposition to the fact—a critical feature of
Einstein’s special theory of relativity—that no signal can travel faster
than the speed of light.
Ironically for Einstein, what was intended to be a potential death
blow to the quantum turned out to be a brilliant prediction when it
was experimentally observed. These experiments revealed that
quantum particles exhibit nonlocal effects. For example, measuring
the polarization of a photon instantaneously determines the spin of
its entangled pair. Simply put, nature is nonlocal. The other question
was whether the spin was real before it was measured—that is,
whether it was determined by what is called a hidden variable of the
system, rather than being a random result of the process of
measurement, as Bohr and his followers believed. According to the
groundbreaking proof of Irish physicist John Bell, this nonlocality in
quantum mechanics has to be the result of the action of a hidden
variable. Interestingly, Bell used the de Broglie–Bohm interpretation,
based on the yet-to-be-observed pilot waves, as an example of a
nonlocal hidden variable theory.
All of this points to a fundamental tension in quantum mechanics.
At one level quantum mechanics is completely deterministic.
According to the Schrödinger equation, if one knows the initial state
of the wave function at some initial time, then one knows exactly
what the state will be at a later time. However, because the wave
function represents a superposition of possible outcomes—it’s a
probability distribution—then if a measurement is made, we will see
only one of such outcomes. In other words, an observation collapses
the function in an indeterministic way. This tension was rigorously
investigated by John von Neumann, one of the great mathematicians
of the century. Von Neumann, along with Eugene Wigner and
Wolfgang Pauli, argued that the measurement problem required a
resolution that would go beyond the current formulation of the theory.
They even went so far as to argue that consciousness itself played a
role in collapsing the wave function. To the contrary, Bohr and his
followers asserted that there was no problem based on
complementarity, and they embraced the inherent contradiction in
quantum mechanics. According to Bohr, there is a sharp divide
between the quantum world and classical measuring device, and the
instantaneous collapse of the wave function is a feature of this
divide. Decades of Nobel Prize–winning physicists have landed on
opposing sides of the measurement problem.
Despite these ongoing and unresolved issues in quantum
mechanics, its formulation as a probabilistic theory where
superpositions can collapse upon measurement has passed a
century of precision experiments and applications. However, as we
shall see, these issues will come back to haunt us when we connect
quantum mechanics to gravity and the early universe.
We’ll get back to that. After all, we’ve still got one more physical
principle to go. But before we do, let’s explore what happens when
we attempt to merge the invariance with the quantum principle.
4

THE ZEN OF QUANTUM FIELDS

Most mornings as a kid I would hear about my mother’s


experiences as a night-shift nurse at Montefiore Hospital in the
Bronx. While she was getting me and my brothers ready for school,
she would tell us about some of her patients and their various
medical predicaments. These stories ignited big questions about my
own mortality, about the fact that my time on this planet was finite
and even uncertain. In hindsight, this was a driving force behind my
decision to study physics, as the field provided a rational approach to
understanding the physical world and our place in it, and would help
make sense of reality and the big questions I had about existence.
As the years went by, my questions about what physics should
say about our place in the universe got overshadowed by the rat
race of publishing articles and landing a permanent faculty position. I
didn’t forget about them, however. In recent years, I have come to
the conclusion that, at least metaphorically, modern physics is
starting to make contact in a non-woo-woo way to some tenets of
Eastern philosophy. Now I can see why Schrödinger and Bohr read
that stuff. Some of these connections have been made in other
works, such as The Tao of Physics. My goal here is to add what I
consider to be a sound new metaphor to this mix of connections,
based on the fusion of invariance and superposition, that is relevant
to the question: Why is there something rather than nothing?
My first introduction to Eastern philosophy was a book entitled
Zen Mind, Beginner’s Mind by Zen master Shunryu Suzuki. Suzuki
had a most interesting metaphor for existence. Imagine a stream
flowing toward a waterfall. When the water leaves the cliff a droplet
of water leaves the stream and at some point, rejoins the stream.
Life is like that droplet of water and the stream is like the universe.
During the time between when we are “born” and when we “pass
away,” we are like that water droplet, feeling like we are separate
from the universe as a whole. But before we are born and after we
die, we are a part of the stream, the universe. I remember wanting
there to be some truth to Suzuki’s metaphor. So, in what way could
we be connected with the universe in the poetic manner described
by Suzuki? According to Zen philosophy this merging can be only
experienced. But I have never experienced Suzuki’s claim, since I
experience having a separate and localized body that’s made up of
matter as I occupy and move through empty space.1 So in place of
experience theoretical physics will have to suffice.
There was a Zen monastery twenty minutes from Brown
University, where I was a graduate student: the Providence Zen
Center. My friend and physics classmate Claire was a regular Zen
practitioner and brought me and a few friends along to engage in a
morning of formal practice. The morning included, among other
forms of meditation, my favorite, breakfast meditation—when eating,
just eat. There were others: after fifteen minutes of chanting words
that I didn’t understand, we sat in silence for three sets of half-hour
sitting meditations followed by walking meditation. This cycle of
sitting and walking meditation would continue for a few hours. Then
the Zen master came out and gave a little talk. She said, “You, me,
this table, this universe are all made from the same fundamental
substance… form is emptiness and emptiness is form.” I later
learned that many Zen practitioners seek to attain a mental state
called satori, where they can experience being one with this
fundamental substance.
How amazing that at the same time I was learning the mother
language of physics called quantum field theory, and the modern
view of physics confirms experimentally what these masters have
experienced subjectively about who we are—necessarily connected
to the universe. But if this insight is true, it must be linked to the
matter that we are made of, atoms and their associated force
carriers. And we will now see that our popular and pedagogical view
of matter has been flawed. To make this clear, let us first look at one
of the fundamental building blocks of matter, the electron.

One of the biggest misconceptions that is nevertheless routinely


taught is that things are made of elementary particles that form the
building blocks of atoms and molecules. It is the case that from a
historical point of view, atoms and a zoo of elementary particles were
discovered first, but when physicists were trying to reconcile the
quantum mechanical nature of these particles with special relativity,
the particle picture was overthrown by a deeper reality of the
quantum field. This field picture of all matter came as a necessity of
trying to reconcile the physics of an electron moving close to the
speed of light in an atom. To properly understand the electron’s
behavior in this relativistic context, we must find a way to merge
special relativity with the electron’s controller, quantum mechanics.
When you do attempt to merge the quantum with relativity, you
immediately see, according to the invariance principle, that the basic
equation of quantum mechanics is not invariant under the space-
time symmetries of special relativity. The main reason is that
ordinary quantum mechanics gives time a preferred status over
space, and in relativity, they are on the same footing. In 1929 English
physicist Paul Dirac ingeniously found a way to unite special
relativity with quantum mechanics.
In hindsight, the hint of this unification comes from
electromagnetism, a classical field theory that Einstein showed is
already invariant under special relativity. The idea of the field was
first introduced by English scientist Michael Faraday to explain his
ingenious experiments involving moving magnets, which he found
would create electric currents in nearby circuits. To explain the action
at a distance between the magnet and the current-carrying wire,
Faraday stated, “I believe that magnetism is actually propagating
itself through this invisible field of influence.” Faraday also said, “I
believe that Electricity has this invisible field of influence and so does
gravity.” His peers rejected the claim about fields as idiotic. This
“invisible field” was considered heretical because the paradigm of
the time was that of a mechanical universe. Similar types of heresy
today would be considered “woo-woo” by critics on YouTube
channels. Biographers say that Faraday died of heartbreak because
his field idea was not accepted during his lifetime. Ironically, long
after Faraday’s rejection from the scientific community,
schoolchildren still play with magnetic filings that trace out magnetic
field lines emanating between the north and south poles of a
magnet.
If you really stop to think of invisible agents acting to impart forces
on objects through empty space, the field concept is actually eerie,
the stuff of witchcraft. It was not until later in the century that the field
idea became accepted. And as we will now see, the field concept will
become a central paradigm underlying modern physics.
James Clerk Maxwell unified a disparate set of electric and
magnetic properties into a coherent framework grounded in the
reality of an electromagnetic field. We can think of the
electromagnetic field as a continuous substance that is distributed
throughout space-time. Electric and magnetic charges will bend the
field in definite ways. Conversely, a warped field can exert forces on
a charged particle. Disturbances of the electromagnetic field can
create waves that move at the speed of light, commonly known as
electromagnetic waves. When these waves vibrate at a frequency on
the order of a trillion cycles per second, we perceive it as visible
light. Einstein earned the Nobel Prize showing that waves of light
can also behave like a quantum particle, the photon, and the wave-
particle duality had to be reckoned with. It took Dirac, German
physicist Pascual Jordan, and Born to realize that the fundamental
substance of the photon is the electromagnetic field that is emitted
when the electron makes quantum jumps from an excited state. In
this case, harmonic vibrations, or quanta, of the electromagnetic
fields gave rise to the photon. If the photon is a particle excitation of
the electromagnetic field, what about the other particles of nature—
are they related to a field?
What Paul Dirac discovered was that to unite special relativity
with quantum mechanics, the invariance principle with the quantum
principle, something radical about the nature of the electron had to
be compromised. He finally figured out, in the spirit of Einstein, how
to make the Schrödinger equation invariant under the symmetries of
special relativity. After succeeding, he found a new symmetry
relating the electron with a mirror electron with negative energy. In
physics when we encounter situations with negative energy, we run
for cover. Negative energy is like falling down a hill with nothing to
stop the fall. But it gets worse than that because all of that energy
will cascade through quantum interactions to the electromagnetic
field, creating an unfathomable explosion that would destroy an
entire galactic system. In Dirac’s case, he used symmetry to
ingeniously reinterpret and repackage this negative energy electron
as a new particle with positive charge and positive energy and called
it a positron, the antiparticle of the electron—a new particle was
born! Two years later the Nobel Prize–winning discovery of the
positron was made by Carl Anderson, confirming quantum field
theory. With the reality of the positron this meant that when an
electron and a positron interact, they annihilate each other and their
masses convert to energy to produce a photon, according to
Einstein’s famous equation that states that matter is equal to energy.
Or an electron and positron can spontaneously be created from a
photon with an energy twice as much as the rest mass of the
electron and positron. This led Dirac and others to no longer think of
the electron particle as a fundamental entity, but a part of one
underlying electron-positron field.
A faithful picture of a quantum field is to imagine the field as the
height of a blanket supported by oscillators (springs). At different
points on the surface of the blanket the spring will vibrate differently.
Now imagine that all these springs are connected to each other.
Then the vibrational pattern of the blanket could be quite
complicated. There are special vibrations of the springs that can
oscillate in sync with each other. Let’s carry the analogy further and
imagine that there is a mass on each spring. Straightforward spring
mechanics tells us that the rate of vibration of the spring, its
frequency, is proportional to the mass. The heavier the spring, the
faster the spring bobs back and forth. So, when the field vibrates at
the frequency equivalent to the mass of the electron, the electron
can be created as a particle from the field’s sympathetic vibration, or
resonance. The particle’s creation from its underlying field is
analogous to Zen master Suzuki’s water droplet that emerges from
the cosmic ocean.2 Likewise, particles are transient entities that
could emerge and return back to their mother field. Where do these
fields live? The electron field is omnipresent and like the
electromagnetic field occupies space-time of the universe. De
Broglie’s hypothesis, which we discussed, is realized because the
vibrational standing wave pattern of the electron’s field can manifest
itself as an electron particle: the particle is nothing but quanta of the
electron’s field vibration.
This leads to a new question about the rest of the matter around
us—could it all be emerging from an excited field? We have a hint
from de Broglie’s hypothesis that relates the mass of a quantum
particle to its wavelength. Amazingly, all matter fields including the
electron share exactly the same building blocks in that they are all
fermionic fields, named after inventor Enrico Fermi. Fermions, the
basic building blocks for all matter, include quarks—the building
blocks of protons and neutrons—neutrinos, and other heavier and
more exotic particles. They all have the key property that they refuse
to occupy the same quantum state. This property is called the
exclusion principle and was discovered by Wolfgang Pauli, who also
discovered the particle called the neutrino. Fermion states are given
by their position and spin. So, if a fermion is at a given position with,
say, spin up, then another fermion with spin up can never occupy the
same position. But because of the exclusion principle, if we continue
to pile up fermions due to their exclusionary behavior, they will form
atoms, molecules, and the various macroscopic forms of matter in
our world.
FIGURE 9: The depiction of the quantum field as a collection of
oscillators that can resonate to create quanta of particles from the field’s
vacuum state.

Here is a story of blackness in physics. The book Atom and


Archetype, which Chris Isham had introduced me to, includes letters
exchanged by Pauli and Jung that reveal that the exclusion principle
presented itself to Pauli in a series of dreams. Pauli had a reputation
for being a harsh critic of other theorists and would swiftly catch
mathematical errors and inconsistencies. He is famously known for
humiliating those with bad ideas by saying, “Your theory is not even
wrong.” In other words, a theory can be incorrect, but at least the
process and intellectual sharpness in developing the theory can still
be valid. Pauli’s statement implies a denigration of the theorist’s
intelligence and process as being laughable. So, it would make
some sense that he kept secret his interactions and dream analysis
with Carl Jung. It’s especially ironic that the kernel of the exclusion
principle came from dreams. Would he have been taken less
seriously or shunned by other colleagues if they knew this?
So, if all matter fields are fermions, what gives them their distinct
flavor? In other words, what makes an electron different from a quark
or a neutrino, given that they are all fermions? The quick answer is
that fermion species can carry different types of charges (electric
charge being one familiar example). We should return to a quick
discussion of fermion field and see what it says about these
differences. Let’s consider the electron field, whose electric charge
sources an electric field that fills space. How does the electric charge
relate to its parent fermion field? It so happens that the electric
charge is beautifully encoded in the fermion field as a special phase
of the field’s vibration. The important idea of phase is central to both
wave and field properties. Imagine two boats rocking up and down
due to the surface-of-ocean waves. If both boats were bobbing at the
same frequency and their high and low points matched, they are said
to be in phase. On the other hand, two identical waves can be out of
phase if their positions do not exactly match up. Here’s an interesting
fact: the phase of the electron field controls the value of its electric
charge. Now there is no reason for the phase of an electron field at
different locations to exactly match, much as rocking boats at
different parts of a tumultuous sea would only be in phase by an
unlikely coincidence. Nevertheless, experiments have revealed that
the value of the electron’s charge is the same regardless of the
location, on Earth, the moon, or anywhere. This means that the
electron field has the same phase, like a couple of miraculous boats.
But what prevents a physics demigod from making the phase of the
electron in New York and Trinidad different, so that Trinidadians get
more electric charge for their dollar? A special form of the principle of
invariance comes to the rescue—this is famously known as local
phase invariance.
Quantum mechanics does allow our demigod to mess with the
phases of fields at different locations, and this will problematize the
fact that we see the same charges everywhere. But it so happens
that there is a magical way to undo the damage the demigod can
make to the phases. If we introduce a photon field to interact with the
electron field the right way, then every time the demigod changes the
phase of the electron, the photon field also makes a compensatory
change in its phase so as to erase the change in the would-be
difference of the electron’s charge. And this magically happens if the
photon field interacts with the electron in a simple way.3 This type of
invariance was invented by the genius Hermann Weyl, and he called
it phase invariance, or more commonly gauge invariance or its
German origin Eichinvarianz. Gauge invariance requires that the
electron field interact with light and guarantees the universality of the
electron’s charge!
Now we can understand what makes the fermion matter fields all
different from each other—it’s their charges, or phase symmetries,
which turn out to be the organizing principle of the standard model of
particle interactions. Once I tell you that the electron has a phase
invariance, you can immediately write down the full electromagnetic
theory, with the unique interaction between the electron and the
photon. And it turns out, the same reasoning goes into determining
the other two nuclear forces, the weak and strong forces, as they are
simply applications of phase symmetries applied to
electromagnetism. The phases of the weak and strong interactions
have more symmetries and give more involved interactions between
the force-carrying particles and the matter fields that carry the weak
and strong charges with which the force-carrying particles interact,
but all these forces strictly come from Einstein’s principle of
invariance, this time applied to the phases of the fermion fields.
Consider electrons in a star millions of light-years away from
Earth. While there are many electron particles, there is only one
electron field that they are born from. Those electrons, and the
electrons that are in you and me, were created as quantum
excitations from the same universal electron field in the early
universe. In fact, every other particle, including quarks and neutrinos,
in the universe are quanta that emerge from their corresponding field
vibrations. It’s in this Suzuki sense that we are tethered to the
fundamental fabric of the universe. All the particles that comprise us
are quantum vibrations of the same quantum fields that extend
across the universe. But does it end here? What is the relationship
between space-time and the quantum fields that permeate it? What
incited these quantum fields to generate the particles that occupy our
universe to turn into stars, planets, and us? And why, since Dirac
demonstrated that antimatter must have been created pairwise with
all that matter, do we not see very much of it around?
Underneath the answers to those questions lies our third
principle: emergence.
5

EMERGENCE

A handful of elements from the periodic table come together to


create living things like you and me. Yet, the elements themselves
are lifeless. How does life emerge from these building blocks? This
question is at the heart of emergence. The first two principles we
discussed, invariance and superposition, combined to give us the
quantum fields we recognize as the building blocks of matter and
subatomic forces. Following Paul Dirac’s prediction of the positron,
new symmetries in nature were discovered as shorter distance
scales, and these new symmetries were experimentally probed with
high-energy particle colliders. And these symmetries also functioned
to specify the nature of the interactions that fields exhibited with
other fields and their associated zoo of particles.
Throughout the last century, physics has been dominated by the
quest to identify the natures of those subatomic forces and a theory
of everything that unifies them. The idea has been that one day we
would be able to find all the fundamental building blocks of the
universe, as well as the rules for their interactions. This approach is
referred to as reductionism, and is essentially motivated by trying to
figure out how to build the universe from the ground up. Superstring
theory is an example of such a unified theory. In string theory, the
basic building blocks are one-dimensional strings. However, even if
we were able to discover such a unified theory, it isn’t clear that it
would be able to explain phenomena such as life or consciousness.
To the contrary: there’s a very good chance that the fundamental
theory could not arise from a reductionist approach.
So we are at a crossroads. On one hand, a major program of
twentieth-century physics took us down the successful road of
reductionism. The symmetries that were discovered informed us
about the patterns and relationships between the elementary forms
of matter and the precise ways that they interacted with each other.
Indeed, the notion that what is fundamental became synonymous
with unveiling more symmetries, at least to some leaders in physics,
and that perspective played a prominent role in my own career. I was
just starting my dissertation work using the exotic symmetries in
superstring theory to solve some of the problems in early universe
cosmology. We know from astrophysical observations that the
universe is expanding. At the earliest stages of the universe’s history,
its environment was exceedingly energetic, hot and dense—
conditions not at all unlike those at a collision of particles in a collider
like the Large Hadron Collider (LHC). So, from the patterns found in
collider experiments, we expect that these symmetries generated by
superstrings were activated in the very early universe, where
superstrings are expected to be the key players. My thesis exported
a special symmetry from string theory into cosmology, called target
space duality, or T-duality, which treats the physics in a large region
of space and a small region of space as being the same. So, as we
approach the big bang singularity, where the universe approached
microscopic distances, we could avoid the big bang singularity with
T-duality. The dream was to use cosmology to test unified theories
like string theory, or other approaches to quantum gravity.
Superstring cosmology is still an important topic in cosmology, and I
still devote some of my research efforts in this direction. So far, the
expectation to unveil more symmetries at the shortest distance
scales works theoretically.
On the other hand, there were clear failures of the reductionist
regime in particle physics. One of the most famous examples of
emergence in quantum physics is superconductivity—really, it’s the
poster child of emergence in physics. In 1911, Nobel laureate Heike
Kamerlingh Onnes observed that when he lowered the temperature
of a metal close to absolute zero, the electric current would flow with
zero resistance. There was no reason to expect how and why the
billions of electrons, which repel each other as well as experiencing
impurities in the metal, should superconduct. After all, lowering the
temperature does not seem to prevent the electrons from repelling
each other or get rid of the impurities in the metal. Many great
physicists, like Einstein, Schrödinger, Heisenberg, Lev Landau, and
Feynman, worked on superconductivity. During this time, many
thought that completely new physics was needed, perhaps a new
law, to account for superconductivity. And it took forty-six years for
the trio of John Bardeen, Leon Cooper, and John Robert Schrieffer
(developers of the “BCS” theory) to show that good old quantum
mechanics and electromagnetism were enough. Superconductivity
doesn’t supplant them. It emerges from them.
The Principle of Emergence: Systems with interacting
elementary constituents can exhibit novel properties that are not
possessed by the constituents themselves.
The reason I am promoting the phenomenon of emergence to a
principle is based on Einstein’s criteria for a principle, because, to
borrow his words, “[A principle is realized by] perceiving in
comprehensive complexes of empirical facts certain general features
which permit of precise formulation.” What are the complexes of
empirical facts in emergence that can transcend its context? There
are many examples of disparate and seemingly unrelated physical
and biological systems where we see emergence. Emergent
behavior also happens in populations of living organisms. Groups of
ants can collectively build a bridge of ants so that others can cross a
barrier of water. The origin of life itself is argued to be an emergent
phenomena. A unicellular bacteria has autonomous properties, such
as motility, replication, and metabolism that its individual molecules,
like proteins, do not exhibit.
But what is at the heart of emergence? How does a system
“know” to exhibit novel collective behavior? These are hard
questions that are currently being researched, and there are some
partial answers. A universal aspect of emergence is the relationships
between the emergent system and the parts that make it up.
Although the emergent property is novel relative to its constituents,
there is an interdependence between the emergence and the
constituents. For example, the system of atoms that gives rise to an
emergent liquid property depends on the collective behavior of the
individual atoms. However, in the solid state, the atoms are on
average located in a repeating array, forming a large-scale crystal. A
simple example of emergence can be seen right in front of your face.
The renowned theoretical physicist Nigel Goldenfeld, who now
directs an institute that focuses on finding links between emergence
in physics and biology, has an experiment that we can all do to
demonstrate an emergent phenomenon. Here is how it goes. First,
push your hand forward in empty space. Second, get a chair and
push the chair until it falls on the floor. The fact that the chair moved
and fell on the floor is emergence at work.
This might sound strange, but it’s true. When you moved your
hand, it was interacting with the air, which is actually a fluid made up
of air molecules. The chair, on the other hand, is a solid. The atoms
of the fluid and the solid obey the same laws of atomic interactions,
yet, despite the sameness of their atomic interactions, the solid state
has new emergent laws of physics. That is, it has new long-range
forces, such as a rigid elastic response from your hand pushing
against it. The origin of these new forces in the solid state arises
from the statistical properties, or collective behavior, of the billions of
atoms. If we work at the level of the description of the solid, we can’t
deduce what the underlying interactions of the atoms are. At the
level of the solid state, all we can deduce is that the continuum
description, rigidity, elasticity, and so on emerges from the collective
behavior of the atoms.
The same goes for superconductivity. What I find interesting in
that story is how BCS hacked superconductivity, not just because it’s
interesting science, but because it gives some guidance for current
problems that we are trying to solve. After all, some of the problems
that we consider to be impossible to solve have been around for a
shorter period than superconductivity.
I had the good fortune to hear some of the story of
superconductivity directly from Leon Cooper, who was my first PhD
adviser before I changed fields to quantum cosmology. During my
first year of graduate school, I didn’t know who Cooper was, and no
one told me that he had won the Nobel Prize. He had the flair of a
Shakespearean actor, wore fine Italian suits, and sported shades on
a well-groomed head of jet black, wavy hair. During our weekly
departmentwide talk, where famous researchers would present their
results, Cooper would sit at the front of the room and ask what
seemed to be naive questions, the type a schoolchild would ask. And
this was exactly the quality of mind that, among other things,
enabled him to access the ingenuous insight that would crack the
nearly fifty-year-old problem of superconductivity.
Superconductivity was couched in the subfield of the physics of
solids, known as solid-state physics.1 One of the leaders of the field,
John Bardeen, who previously shared a Nobel Prize for the
discovery of the transistor, had been tirelessly working on
superconductivity for years with no luck. Bardeen was well aware of
the decades of failed attempts to explain superconductivity, and he
decided he had to get an outsider’s perspective and tools to bring
new life to the problem. So, he sought out a theorist that had a fresh
pair of eyes and wasn’t jaded by the biases that may be formed by
experts in the field.
I remember the serendipitous early summer day I was driving
from Hanover, New Hampshire, to New York City and decided to
stop by Brown with the hopes of seeing Professor Cooper. It had
been almost a decade since I had seen him in person. I caught wind
that he was retiring (but still fully engaged in research, as he still is).
During summers, many faculty are traveling to conferences, but
when I got to the physics building at Brown, Cooper was there in his
spacious office, filled with books and covered with blackboards,
doing a calculation. We sat down for a chat. He immediately asked
me what I was working on.
I thought that I was going to impress him with my new take on the
matter-antimatter asymmetry in the universe. Cooper stopped me
and said, “You should find a real problem and solve it. Many people
put their hands up in surrender when a problem gets too hard and
claim it’s impossible.” I took this to be both a challenge and
validating. What might at face value look like Cooper’s rejection of
my idea I took instead as him holding me to a higher standard and
the expectation to solve a big problem. Even today, I try to live up to
my former adviser’s version of tough love with my students—to
recognize and help awaken their hidden talents. Up till that point, I
was playing it safe and avoiding physics problems that I thought only
the most able of physicists should have the permission to work on. I
asked Cooper, “How did you solve superconductivity?” What he said
gave me some strategies for approaching my own problems.
Cooper went on to tell me that he was trained in another field,
theoretical particle physics, and had mastery of the techniques new
to that field, such as Feynman diagrams. When he was invited to
work with John Bardeen in condensed-matter physics, as the field
working on superconductivity was known, he had an unsullied and
less-biased take on the nature of the problems those physicists were
facing. For one, as we’ve seen in this book, particle physics
concerned itself with discovering the nature of subatomic forces by
exploiting the quantum scattering processes between elementary
and nuclear particles. Solid state environments concern the behavior
of billions of interacting electrons in an environment filled with other
atoms usually organized in the form of a periodic crystal lattice.
Superconductivity was both a conceptual and mathematically
technical dragon to slay. One major obstacle was that the problem
seemed to require solving the Schrödinger equation for a wave
function of billions of electrons interacting with a lattice of metallic
atoms—the many-body wave function. No one working on the
problem, no matter how technically skilled, has been able to
surmount the mathematically crippling wall of solving the many-body
wave function. According to Cooper, Bardeen “omitted to mention
that practically every famous physicist of the 20th century had
worked on the problem and failed.”2
Cooper quickly encountered the daunting and insurmountable
equations. On a seventeen-hour trip to New York City, he tirelessly
tried his extensive bag of mathematical tricks but got nowhere. He
ran out of technical steam and started feeling that the equations
were preventing him from seeing the root of the problem. So he
decided to step back from the equations and think intuitively about
the problem. And then he made a simple and ingenious guess. Part
of his mental wizardry was to simplify the problem and avoid
unnecessary details, decisions geared toward making the problem
more tractable.
As a hint into Cooper’s insight, recall that electrons carry a tiny
magnetic pole. This pole can also obstruct their motion, say when
they are flowing in a current, due to the magnetic forces of
surrounding electrons, which cause deflection and electrical
resistance. It seemed that, if superconductivity were going to be
possible, then the golden rule that electrons necessarily repelled
each other had to be broken. Cooper realized that if the electrons
could pair up, with their spins oppositely aligned, then the members
of each pair would lose their identity as electrons, and their overall
spin would vanish, mitigating the local resistance. An emergent
phenomenon, the Cooper pair, was born. But the grouping doesn’t
stop there. When all the electrons pair up, they clump together to
collectively behave as one object and move in a ghostly fashion
through obstructions in the metal. Cooper likened it to a line of ice-
skaters, arm in arm: “If one skater hits a bump, she is supported by
all the other skaters moving along with [the line].” In other emergent
phenomena in condensed matter, this long-range order is a
collective behavior of the individual electrons or atoms.
The formation of Cooper pairs led to a handful of other emergent
properties in the superconductor. First, in order for the supercurrent
to maintain itself, the superconducting environment would have to
expel any magnetic field. This observation is known as the Meissner
effect and is predicted by the BCS theory—it is the reason magnets
levitate above superconductors. The underlying physics in
superconductivity was later found in other systems, such as neutron
stars. The extremely dense environment of neutron stars enables
neutrons to Cooper pair and exhibit the collective behavior of a
superconducting fluid, called a superfluid. Another Nobel Prize was
awarded to Yoichiro Nambu, who applied the BCS theory to
understand the emergence of particles called the pion, which was
found to be a Cooper pair of quarks. BCS theory also inspired some
of the architects of the standard model of particle physics to think
about how mass could emerge from a similar type of symmetry
breaking, and we will discuss this in an upcoming chapter.
In a seminal essay entitled “More Is Different,” Nobel laureate
Philip Anderson puts emergence at center stage over reductionism
in physics: “The ability to reduce everything to simple fundamental
laws does not imply the ability to start from those laws and
reconstruct the universe.… At each stage of [emergence] entirely
new laws, concepts, and generalizations are necessary.” Anderson
goes on to identify the organizing principle that is behind most
emergence in condensed-matter systems, that is, symmetry
breaking. When we see symmetries, we often see an underlying
pattern of phenomena. For example, in relativity, the space-time
symmetry inherent in the laws of motion functions to give relative
lengths and time for different moving observers. Emmy Noether
proved that symmetries are linked with conservation laws. And
symmetry breaking signals new properties that are hidden from the
symmetric realm.
To see this, consider a piece of metal, like iron. Like a
checkerboard is a repeating arrangement of black and red squares,
all metals are repeating periodic arrangements of atoms with
electrons waiting to easily flow from one site to the other. At each
atomic site the electrons carry a tiny magnetic pole due to their
quantum spin. Recall that the quantum spin can either be up or
down, reflecting its quantum nature. At high temperatures the billions
of the electron spins point in a random direction. If we add all the
spins, because of the random orientations, the total spin of the
system cancels out to zero. Since the total spin and magnetization is
zero there is no preferred direction of all the spins and the system
has a symmetry that is invariant under rotations. This means that the
iron has a symmetry that is analogous to the symmetry of a sphere.
However, as we lower the temperature to a critical value, the spins
all on average spontaneously pick a direction, attaining a net
magnetic field. This is like balancing a pencil on its tip. As a result,
the spherical symmetry is broken by the emergence of magnetism.
Not only is magnetism emergent, but the preferred direction that
breaks the symmetry creates rigid directions that can support the
transmission of magnetic waves. In all the examples in condensed-
matter physics, the constituents, such as electrons or other fermions,
exhibit cooperative behavior similar to the swarm of ants or bees that
collectively perform a task that an individual member cannot
accomplish alone.

FIGURE 10: An example of spontaneous symmetry breaking. In image A


the forces acting on the pencil have a rotational symmetry. However, this
system is unstable because the gravitational force in the z-direction will
break the symmetry. In image B the pencil picks a direction, which results
in breaking the original rotational symmetry. Any direction could have
been selected, and this randomness is a sign of the symmetry being
spontaneously broken.

In hindsight it was also discovered that superconductivity


emerged from the breaking of another type of symmetry, not related
to space-time but closer to those found in the fundamental
interactions, such as the strong and weak nuclear forces. The
principle of emergence and symmetry breaking turns out to be at the
heart of important matters in the fundamental forces. As we will
explore in the next chapter, the nuclear interactions, such as the
weak and strong forces, are governed by symmetries associated
with the charges of elementary fields and particles. And the breaking
of these symmetries also has essential properties, such as the origin
of mass and the emergence of matter over antimatter in our
universe. Even stranger is the idea that space and time are also
emergent properties. By analogy, these are atoms of space and time
whose collective behavior can give rise to the malleable space-time
fabric that Einstein discovered.
There has always been a deep interplay between physics and
other scientific disciplines like biology, chemistry, and the social
sciences, where even more mysterious forms of emergence occur.
Does the organizing principle of symmetry breaking seen in physical
systems apply to understanding emergence in other domains? Are
there other organizing principles that go beyond symmetry breaking?
And maybe even beyond physics?
6

IF BASQUIAT WERE A PHYSICIST

The wisdom of John Bardeen, a two-time Nobel laureate, to seek


and integrate the outsiders’ perspective of Leon Cooper not only
enabled cracking the code of superconductivity but opened the
floodgate for breakthroughs in other branches of the physical
sciences and technology. If we want to catalyze more solutions to
the current mysteries we face, we could try to replicate these
examples of scientific inclusivity that Bardeen and others
exemplified. These days there is a big push for and rhetoric
surrounding diversity and inclusivity in science. But science has
fallen short in benefiting from effectively embracing outsiders, or
even wanting to do so. Science is carried out by individuals and
groups of individuals, and the principle of emergence also acts on
scientific societies sustaining forces that act to prevent the full
benefits from the contributions and presence of outsiders. In this
chapter, we explore through the lens of the science of groups of
people—the science of sociology—to see how the scientific
community and individuals can inspire and enhance more
innovations and innovators. To gain insight into these issues, I
present two stories.

SCENE 1
Eleven years ago, Jim Gates, a theoretical physicist with a Frederick
Douglass–style ’fro, was the chair of Howard University’s physics
department. I considered going to Howard as an undergraduate; it is
a prominent, historically Black university with famous alums like
Thurgood Marshall, but I opted instead for a private Quaker college.
During that time, Howard University made a bold move by poaching
a handful of Black physics professors from prominent, predominantly
white institutions, and this inspired Gates to move from the
University of Maryland. During his time at Howard University, Gates
and his colleague Hitoshi Nishino were researching a special
symmetry called supersymmetry. Supersymmetry is a theory that
attempts to connect the fermions, such as electrons, of our world to
the force carriers, such as photons. Although in our experience we
see both these fermions and bosons, as the force carriers are called,
Gates and Nishino were thinking about a class of particles that we
don’t typically experience, known as anyons. Anyons only exist in
two-dimensional systems, and they emerge from matter particles,
but they enable the violation of a fundamental law that matter can
only have integer or half-integer spins. Gates and Nishino’s work
was aimed at developing a theory that linked the supersymmetric
relationship between fermions and bosons to anyons. Recall that all
observed matter has either integer or half-integer quantum spin.1 For
example, the electron, a main building block for molecules, has half-
integer spin. Gates and Nishino wrote a beautiful set of equations
that made these anyons supersymmetric and discovered a surprising
feature that the equations were conformally invariant—we won’t go
into the details here, but it’s a symmetry that related the microcosm
to the macrocosm.2 The result was published in a respectable
physics journal and enjoyed a modest number of citations.
In the subsequent eleven years, a theory developed from string
theory called M-theory, which attempts to unify all forces, underwent
a conceptual revolution—it encoded a property called holography. A
holographic theory is one where gravity is encoded in another theory,
operating in one fewer dimension, without gravity. For example,
gravitational physics in our three spatial dimensions could be
holographically encoded in a two-dimensional theory with no gravity.
This work was famously christened the ABJM theory, named after
the authors, and was, and is, considered one of the most important
results in theoretical physics in decades. The four authors much later
discovered that the exact supersymmetric and conformal invariant
equations that Gates and Nishino derived had a holographic
description in M-theory. Many physicists, including me, were not, and
still are not, aware of the original equations of Gates and Nishino. To
be fair, the authors of ABJM actually cited the Gates-Nishino work,
yet it still did not rise to the acclaim that it may have deserved. After
all, the equations that ABJM used were the same that Gates and
Nishino had derived eleven years earlier. Why did the community not
call this theory Gates-Nishino-ABJM? Why did the community not
notice the importance of the original Gates-Nishino work?

SCENE 2
After a few months into my second postdoc, I stopped going to my
office to work. The dozen or so postdocs in the theory group were
very interactive. However, time after time, I found my attempts to
interact with my peers were not reciprocated, and were even
ignored. One day a good friend, Brian Keating, who was a postdoc at
Caltech, was visiting our group. Brian, who is white, pulled me aside
and said, “I know what’s going on. I know why you’re not coming to
your office. I overheard a conversation with some other postdocs,
and they said that they want to punish you.” So, what did I do to
them that would warrant punishment by shunning me? My friend
volunteered the reason: “They feel that they had to work so hard to
get to the top and you got in easily, through affirmative action.” I must
admit harboring both disdain for and envy of my postdoctoral
colleagues. Most of them grew up with privilege that I did not have
and a sense of entitlement that the enterprise of science belonged to
people like them. I also presumed that their relationship with physics
was different from mine.
For me physics was literally a tool for survival. Reading physics
books and solving problems kept me away from the streets of the
Bronx. Physics paid my way into college and graduate school, and
unlocked the shackles of a likely life of poverty in the Bronx. There
were times when I would stay up all night playing with physics
problems and equations. Yet, I remain thankful to Brian, because,
after that day, I made sure that most of my publications and research
over the next three years as a postdoc were independently
administered and authored. I did not want to be further penalized by
colleagues and peers since I could not change their perception of my
not deserving admittance to their elite club. By doing independent
work, I felt that it would address the perception of whatever
shortcomings may inhibit my future employment. While it was useful
to learn how to complete independent work, my strategy still did not
erase the perception and treatment that I would receive from
colleagues throughout my career.
Because I lacked a feeling of belonging in the group, and so that I
could continue to be productive, I moved most of my calculational
operations to a café across from Stanford’s computer science
department and I worked by myself. In hindsight, this isolation was a
blessing in disguise; it forced me to develop my outside-the-box
thinking. At the time, the discovery of dark energy—a concept
closely related to a parameter in the general theory of relativity
known as the cosmological constant—prompted the entire group,
including me, to rethink the very foundations of theoretical physics.
And because I had just spent two productive years in London at
Imperial College, where I had developed an improvisational and
visual style of approaching problems coupled with Jungian dream
analysis that Chris Isham had trained me in, I was very prepared for
some outside-the-box thinking about the issue of dark energy. Once I
had developed a clear idea of what I wanted to pursue with those
techniques, then I would unleash an arsenal of traditionally
mathematical devices that I mastered to develop a model. I naturally
kept this atypical research strategy to myself in fear of further stigma.
One morning after forcing myself to go to campus, I was struck by
an insight that had come in one of those dreams I had told Isham
about a year previously. The dream inspired me to focus on the
particular properties of space-time called discrete transformations.
You can see an example of a discrete transformation when you look
in a mirror, which makes your left hand look like it is the right hand of
your reflection. I had a sense that I was onto something important,
so at my café office, and with a celebratory beer, I started doodling
inchoate diagrams on a napkin. I had developed this strategy of free
play by presenting many competing rough sketches from my mind’s
eye of the physics I was developing, before committing to any
mathematics. At that stage of my investigations, there were no
equations, just different sketches that resembled Picasso-like
drawings of the physics of accelerating space-times and their
discrete transformations—their reflections.
One day not very long after, I noticed the head of our theoretical
physics group, Michael Peskin, on a physics chat stroll with the
golden-child postdoc of our group, the guy that all the postdocs
wanted to be like. As Peskin and the golden boy walked by, I couldn’t
hold myself back. I said nervously, “Michael, I think I got something
interesting about some work on the cosmological constant.” Peskin
engaged me, and I started blurting out what in hindsight were pretty
zany ideas. The golden child smirked and said with a tone of
dismissal, “You and your crazy ideas again.” In fact, golden boy was
one of those whose rejection kept me from hanging out in the theory
center. But Peskin saw something and said, “That’s interesting, why
don’t you come to my office next week and tell me more about it.”
I did that. When Peskin takes a physicist seriously, he usually
throws a question back at them, like a physics koan. After I
explained my inchoate idea, Peskin challenged me to do some
difficult warm-up calculation that would help my idea take shape. For
me, it was important that he took me seriously and challenged me.
Over eleven grueling months, with much toil and many calculations,
those doodles on a napkin transmuted into a publication in the top
journal in physics. We had created a new approach to understanding
how matter over antimatter was created in the early universe. The
paper opened up new directions in astrophysics and cosmology
research; it has been cited over two hundred times.
Michael Peskin is regarded as the “oracle” among theoretical
physicists; he is very much an insider. But Peskin is unusual; he
possesses a quality that enabled him to appreciate and perceive
value in my style of doing physics and in the ideas that I generated
that were different from the norm.
Coincidentally, Peskin was an office mate at Harvard in the late
1970s with Jim Gates. At the time, Gates was doing pioneering yet
nonconventional work on combining supersymmetry with gravity.
Peskin was a young prodigy working in a different field of particle
physics. Gates shared with me that Peskin took his work so seriously
that he spent ample time learning the daunting mathematics behind
supersymmetry and even applied it to his work.
Were the expectations and consequences for doing innovative
and transformational research simply lower for Gates, me, and other
minorities in science? In April 2020 collaborators from the computer,
education, and linguistics departments at Stanford University
published results utilizing machine learning that asked the following
question: Do women and minority scientists innovate as much as
their white male counterparts? After a careful look at the research
output and impact of 1.2 million women and minority scientists over
twenty years, the findings concluded that women and minorities
innovate novel contributions more than the majority group.
Innovation involves publishing novel contributions that are used often
by other researchers. Novelty enables new connections between
ideas that generate knowledge. If those groups innovate more, how
do we explain their lack of prominence and promotion in the scientific
enterprise? In their paper, entitled The Diversity Innovation Paradox,
the Stanford researchers explained that “Novel contributions by
gender and racial minorities are taken up by other scholars at lower
rates.… There may be unwarranted reproduction of stratification in
academic careers that discounts diversity’s role in innovation.”
One of the authors of the research article was asked why the
innovation from minoritized groups went unnoticed; he responded
that “the fresh perspectives that women and nonwhite scholars bring
are atypical and can sometimes be hard to grasp, so they get
devalued by the majority.” A scientist like Peskin is rare: he could
see talent when other insiders saw African Americans as deficient
and as interlopers.
In what follows, drawing on the two stories I just presented, we
will dive deeper into how science and scientists can be better
positioned to enable future breakthroughs that would otherwise go
unnoticed or disabled. Let us keep in mind that there are many
stories like this. We will end with an observation of what the scientific
community can learn from the art world to transform itself to do
better science.
In a seminal work social theorist Robert Merton observed that
better-known scientists get more recognition than a lesser-known
scientist for the same achievement—he called it the Matthew effect,
which alludes to a biblical saying, “for unto every one that hath shall
be given, and he shall have abundance: but from him that hath not
shall be taken away even that which he hath” (Matthew 25:29). This
is also consistent with the well-known saying, “The rich get richer
and the poor get poorer.” It can be tempting to explain away Gates
and Nishino’s predicament with the Matthew effect. According to the
Matthew effect, Gates’s paper would have likely been more widely
cited and recognized had he been the chair at the top physics
department, like Princeton, which was the home institution to one of
the authors of the ABJM theory. This is partially true, because when
he published the work with Nishino, Gates was stationed at a
historically Black university that was trying to upgrade its visibility to
the physics community. However, the Matthew effect does not fully
explain why Gates did not get the credit eleven years later, since he
is currently well known in the scientific community: he is the
president of the American Physics Society, a member of the National
Academy of Sciences, and a winner of the highest award given to a
scientist by the U.S government, the National Medal of Science.
Insight into Gates’s lack of recognition is tied with my experiences
as a young scientist in spaces of high repute, where I was
stigmatized and shunned because of social presumptions about my
belonging in their cohort. And as we will see in what follows, there is
a hidden gem for the advancement of science provided that we are
able to clearly perceive the functioning of the blind spots of the
scientific hegemony and place value on members who have
developed new ways of innovating. To do this, I invite you to delve
into the hidden ways psychosociological phenomena impact science.

The various activities of scientific enterprise are carried out by a


community of scientists, who form a scientific social structure—a
normative order. What are the consequences for science when
scientists act within social structures? We know that the
institutionalization of science as academic disciplines facilitated its
growth and its accomplishments.3 Is there an underappreciated flip
side to this? Can the social order of science generate blind spots
and even enable bad faith that prevent a better understanding of the
mysteries we ponder? To answer these questions, it is useful to
develop some tools of analysis from social theory.

SOCIAL NORMS, CULTURAL NORMS


While some might not have access to the kinds of experience of a
Black person in America that would give the intuition of W. E. B. du
Bois’s notion of “double consciousness,” social theory provides some
insights and tools to enable us to transcend our collective blind spots
to the benefit of scientific progress.4 My curiosity about my
predicament as a scientist and desire to break new scientific ground
led me to the works of Émile Durkheim, one of the founders of
sociology. Durkheim observed that we all live within social and
cultural orders, that our lives are regulated by shared social values
and moral obligations that he referred to as the conscience
collective,5 and by shared cultural norms, which constitute meaning,
and our commonsensical understanding.6 Social values validate
social norms, which enable expectations that are differentiated
according to social position, while cultural norms regulate the nature
of meaningful actions. Social norms distinguish between right and
wrong, while cultural norms constitute the difference between sense
and nonsense, being an insider or an outsider.
When thinking about the normative order of science, we need to
distinguish between two sets of norms, cultural and social. The first
involves the culture that regulates scientific activity, for example, that
theories must be logically coherent and empirically warrantable.
These norms differentiate between science and nonscience, as with
paleontology and creationism, or cosmology and flat-earth beliefs.
While this differentiation is permeable around the edges, the
distinction between science and nonscience must be sustained.
Social norms are the normative orders that create expectations
within specific social circles of scientists; they define the boundary of
what is accepted and valued within a specific group of scientists. For
example, two competing yet viable scientific theories could vary in
validity within different scientific groups on grounds of “taste” or the
reputation of the architects of the theory. These judgments are
mostly subjective but have consequences that help explain the two
stories presented earlier.

DEVIANCE BOTH ACTIVE AND PASSIVE


We have talked about social and cultural norms and how they create
the boundary for actions that are considered acceptable. But what
about those who violate those expectations? Well, there’s a word for
such violators: deviant. Durkheim posited that the social order is
maintained and replicated by the existence of deviance; you cannot
have one without the other. And this has big implications for
innovation in science. In the first instance, violations of cultural
norms make no sense, so we endeavor to make sense of them,
often by labeling the violation, or violator, as “crazy.” Violations of
social norms are “wrong,” and to reinforce our sense of what is right,
we negatively sanction known violations of social norms. Thus,
social and cultural orders are maintained by penalizing deviant
behavior. When social and cultural deviance are punished, the
punishment constitutes or reinforces the boundary between what is
allowed and what is disallowed.7
We need to be careful about the use of the word deviance,
because it actually has two meanings. One meaning, the usual one,
connotes a deliberate violation and disruption of the laws. You might
think of someone who robs a grocery store. We can call this active
deviance. However, the current social order in the scientific
community institutionalizes and implicitly promulgates that people of
color or women as categories of persons are incapable of doing
science as well as white men. You might think of the ways such
stigma has followed me through every level of study for reasons that
have nothing to do with my actual ability, or the ways that I have
described being pushed out of social groups for the crime of existing
in the community at all. I am deviant by default. We call it passive
deviance. The persistent unwelcoming behavior a scientific
community exerts on minorities engenders isolation, which in turn
can stymie productivity. The irony here is that passive deviant actors
can be a positive asset to the scientific social order; this is, at least in
part, because they are more likely to be positive deviants, to violate
social and cultural conventions that restrict the bounds of scientific
creativity. This is consistent with the research findings in the
Diversity Innovation Paradox article.
Deviant violations are punished; however, whatever form the
punishment might take, this will both discourage future violations and
reinforce institutionalized cultural and social expectations. As a
consequence, physicists might be motivated to avoid forms of
deviance that could result in creative breakthroughs. It is important to
distinguish between deviant behavior that is harmful and deviant
behavior that results in scientific innovations. Let us explore whether
the latter may emerge among minority scientists who carry markers
over which we have no control.
Marginal people in disciplines like physics may be in a valuable
position to innovate fundamentally because they are likely to expand
the plurality of ideas, approaches, and techniques in the discipline.8
They are less likely than those who “fit in” to feel the pressure to
remain within the constraints of their discipline. In my case, though I
had the same technical training as my postdoc peers, my social
isolation from the group enabled me to both not replicate conceptual
blind spots and to embrace ideas on the fringes of established
knowledge. But how could science emancipate itself from this fate of
suppressing contributions from outsiders?

As a youngster in the Bronx, I lived two blocks away from the last
stop on the 2 train, which served as a depot for trains to be serviced.
During the sleeping hours, regardless of the weather or unseen
dangers, graffiti artists would gather at the depot to work on their
masterpieces; my favorite was the larger-than-life spray-paint portrait
of the Marvel comic archenemy of the Fantastic Four, Dr. Doom,
which covered an entire train car. The 2-train depot served as the
local art gallery for schoolchildren as we waited for the daily yellow
school bus. I and many of my friends were enchanted by comics and
drew our own characters; we looked up to the enigmatic graffiti
artists as intrepid heroes, modern-day Peter Pans who broke the
rules to do their art.
Painting graffiti on the subway came with the penalty of a fine or
criminal arrest. Graffiti expression walks a fine line between art and
vandalism. If a graffiti artist does not have permission, then the art is
deemed illegal and is considered vandalism. Andy Warhol, an insider
in the New York art scene, gave graffiti and street artists permission
and validation to have their graffiti integrated into the art
establishment. In the 1990s Brazilian graffiti artists were harassed
and sometimes shot at by the police. Ironically, today Brazilian graffiti
has led to founding art schools in low-income neighborhoods and to
a collaboration with police to paint murals in devastated areas.
These days, you can find graffiti art in the most prestigious art
museums and galleries on the planet; the “best” work sells for
astronomical amounts.
There are some valuable lessons that science can learn from
graffiti. The graffiti artists benefited the art establishment when they
embraced their outsider status and continued making graffiti
independent from the hallowed art galleries. For example, post-
graffiti artists like Banksy, Samo (a graffiti duo of which Basquiat was
a part), and Keith Haring established themselves, and continued to
work on the outskirts, while gaining mainstream acclaim. Renowned
graffiti writer Eric Felisbret says it well:

From the perspective of a graffiti artist, the debate about


whether graffiti is art or crime is pointless because, ideally, it is
both. In the graffiti community a writer cannot achieve status
solely based on artistic ability. The writer must also be willing
to work outside the law and assume great risk. The movement
—which I have been documenting in New York for over 30
years—was founded on this principle and it defines its
essence.9

Outsiders who craft nonconventional ideas and develop new


techniques, similar to graffiti, can be seen as vandals, and that
“vandalism” may be penalized. Innovating outside the mainstream is
hugely risky. However, the realization that some forms of deviance
result in positive accomplishments was a game changer for me. The
sense of alienation I felt in science, with all its rejection and stigma,
also comes with the advantages of being an outsider. There is value
in having an outsider’s perspective and an opportunity for innovation
from being in my natural state when I am taking intellectual risks. In
2018 I wrote in an essay:

I’ve come to realize that when you fit in, you might have to
worry about maintaining your place in the proverbial club.
There are penalties for going elsewhere or doing things your
own way, as nonconformity can feel threatening to the others
in your circle.
So, I eventually became comfortable being the outsider.
And since I was never an insider, I didn’t have to worry that
colleagues might laugh at me for an unlikely approach. Many
times, that unorthodox approach actually led to new
understandings.

Both the art world and the physics world have deviant actors. The
difference is that the art world has embraced graffiti, while the
scientific community has yet to embrace those who take risks. It is
clear that embracing graffiti was good for the graffiti artists, but I’d
argue that it was equally good for the conventional art world. Imagine
contemporary art without Basquiat’s beautiful and unsettling figures,
or Banksy’s consistent challenges to authority in both art and politics.
Imagine city streets in which all the street art and the brick-side
murals are painted over. I don’t like to imagine a world without the
richness and the beauty that those contributions brought.
I like to imagine Basquiat as a physicist. I think of him strolling
down his university’s hallways, maybe stopping briefly to chat with
his colleagues. I think that the blackboards in his offices would be
covered with drawings as well as equations. I think that students
would come to peek at the art while he worked. If Basquiat were a
physicist, his work would be as unconventional as his office. I expect
that he would break the rules, and as graffiti artists do, he would take
pleasure in doing so. And those students who came to watch him
working, I think that they’d learn how to break the rules too. If
Basquiat were a physicist, he would be able to recognize
immediately the value of contributions that others in the field might
see as simple heresy. And I’d like to think that a physics community
that welcomed Basquiat into its fold would want to take the risks
associated with giving such contributions a fair shot, even if the rest
of the community might not immediately see the value in Basquiat’s
work. I think such a community would be richer for his presence,
and, for his presence, able to grow richer over time.
A naive conversation about diversity in sciences, often filled with
gesticulations of identity politics, sustains the smoke cloud that
obscures the real issue, the true value of not completely belonging,
of not always being comfortable around others, a discomfort that
difference brings. Assuming that outsiders attain the competencies
of the field, their outside perspective and nonconformity can be
exactly what is needed to facilitate major breakthroughs. In fact,
many significant innovations in science came from someone who
was an outsider in a given field, someone who applied a new
technique or perspective from another field. Perhaps this was
because they were valued within both disciplines. Perhaps it is time
to value and elevate minorities, thus enabling them to make major
contributions, not in spite of their outsider’s perspective, but because
of it.
PART II
COSMIC IMPROVISATIONS

Cosmic Improvisations’ narrative and structure will proceed like a


real-time jazz improvisation, wherein readers and I will solo on some
of the most pressing issues and controversies in fundamental
physics and cosmology. To this end, I have structured the rest of the
book like a jazz album. In the jazz tradition, a tune often begins with
the “head,” the main melodic theme. The head is connected to the
harmonic and rhythmic structure—the form—of the piece. A soloist
improvises over this form, using the head as a context in which to
search for and discover new melodic ideas. Saxophonist John
Coltrane’s improvisations exemplify this process: he integrated and
creolized musical forms that were thought to be incompatible—as in,
for instance, his masterpiece, A Love Supreme. He broke with
tradition as he transcended it, incorporating in it the new musical
styles he created. Using my three principles in Part I of the book,
we’ll explore how new ideas in physics are created, and try to
answer some of the biggest puzzles in cosmology—including the
nature of the big bang, the cosmic origin of life, and the role of
consciousness in the universe, as well as discussing the possible
quantum nature of gravity, and the role of dark matter and dark
energy in cosmic evolution. And we’ll do this drawing on my
improvisations and work, as well as those I’ve undertaken with my
collaborators. But we’ll also look at the work of other cosmologists
and scientists. While improvisational logic is key to guiding us to new
landscapes in physics, it’s not enough; as every jazz musician
knows, we must also learn from the improvisations of others.
7

WHAT BANGED?

The universe is expanding. Why should we care? After all, it seems


like the expansion has no effect on us, right? We are bound by
gravity to our solar system, and if you look out at the night sky, you
see stars, no different from our own sun. Like grains of sand on a
beach, our sun is as typical as the other stars in our galaxy, filled
with hundreds of billions of suns. But the expansion of the universe
matters to those stars, which means it matters to us. Stars are
necessary not only for sustaining life as we know it with their light
and heat, but for functioning as the manufacturer of the material of
planets and life. As we will soon see, there is a startling connection
between the early evolution of the universe and the creation of the
substances necessary to form stars, which points to a delicate
interplay between gravity and quantum physics. And some of this
interplay remains unsolved.
Einstein taught us that space and time are more than the location
of an event, but that, like electromagnetism, space-time itself is a
dynamical—in its case, gravitational—field. Not only does matter live
on space-time, but space-time lives on matter. Our most direct
experience of this fact is that the sun warps the gravitational field,
like a sitting person’s weight warps a cushion, bending space-time
and creating orbital contours through which planets can move. In
general relativity the space-time field, called the metric tensor,
encodes information about the curvature of space-time in the
presence of matter and energy. As John Wheeler famously states,
“Matter tells spacetime how to bend and spacetime tells matter how
to move.” But space-time can dance in many different ways
depending on the configuration of matter and energy that interact
with the gravitational field. A famous example is the spherically
symmetric space-time of a black hole that is sourced by a collapsed
star. The mass density is so high that space warps such that the
emitted light cannot escape the highly curved space-time.
By now it is popular knowledge that the universe is expanding,
and this has been experimentally confirmed by looking at receding
galaxies and racing away exploding stars called supernovae. When
the distribution of matter and energy is the same in every direction
and at every point in the universe, the equations of general relativity
predict that the universe’s space-time will expand. A simple way to
intuit this type of physics is to imagine the surface of an expanding
balloon with points fixed on the surface. Imagine that all points on
the balloon’s surface is the environment of a galaxy.
As the balloon expands, the galaxies appear to recede from each
other. An observer in a given galaxy is fixed at the same point on the
balloon’s surface despite its expansion. What’s moving is the space
(rubber) between the points, or galaxies, on the balloon. If you can
imagine that a region on the surface of the balloon is a region of
three-dimensional space, then the analogy comes pretty close to
illustrating the dynamics of an expanding universe. Where does the
analogy break down? The rubber makes up the balloon, but space in
our universe seems empty, yet dynamical. It appears that space is
continually being created in an expanding universe, and we will
discuss this perplexing fact in a later chapter.
But what is the significance of the expanding universe, aside from
making our universe extremely large to house billions of galaxies?
Even after I wrote my first research paper that involved some new
features of the expanding universe, I wondered about that question,
but was too embarrassed to mention my puzzlement to others. I
soon discovered that a copioneer of the expanding solution of
general relativity was a Belgian Catholic priest and theoretical
physicist, Georges Lemaître, in 1927, who was driven to reconcile
his religious beliefs with his love of and conviction for the veracity of
general relativity.1 From this solution Lemaître predicted that the
expansion would make galaxies recede from each other at a speed
that is proportional to the distance between them. Two years later
Edwin Hubble confirmed this prediction by showing with telescopic
data that the light emitted from galaxies was redshifted as predicted
by Lemaître’s expanding solution—indicating that galaxies are
moving away from each other and us.
Lemaître seemed to have been able to reconcile his religious
conviction with the materialist explanation from relativity theory by
positing that the universe emerged from an initial point of “creation”
where all the matter in the universe was concentrated. He called this
point the “primeval atom” or the “cosmic egg, exploding from the
moment of creation.” This view of the creation of the universe is also
found in many proto-Indo-European and African cultures. This later
became known as the big bang, coined by English astronomer Fred
Hoyle, who was a proponent of a competing theory coined the
steady-state universe. Lemaître says it quite poetically: “We can
compare space-time to an open conic cup.… The bottom of the cup
is the origin of atomic disintegration; it is the first instant at the
bottom of space-time, the now which has no yesterday because,
yesterday, there has no space.”
According to Lemaître’s solution, distances increase as time in
the universe elapses. If we run the cosmic clock backward then there
will be a time in the past when all distances tend to zero, leading to
the nonexistence of space. The geometry of this expanding universe
can be recast in a mathematic form that the universe’s beginning
looks like a point in a four-dimensional surface that spreads into a
cone in the future. The priest intuited that the entire content of the
universe was contained at this point of origin, the “bottom of the cup”
that he called the “primeval atom” that burst into an expanding cone.
However, he did not provide a mechanism for this disintegration of
the primeval atom, nor what conditions led to this initial state of the
universe. And there is another serious problem associated with this
state of infinite density. As the universe shrinks, the curvature tends
to infinity. When physical quantities of a theory go to infinity, we call it
a singularity and it usually means that the theory itself breaks down
and can’t be trusted. But there are some subtleties about cosmic
singularities that deserve more attention and background, which we
will discuss in a later chapter. History has shown that singularities
signal new physics that may resolve the singularity. But there is
another clue that may give us insight about the very birth of our
universe’s space-time and what may have sourced it.
When the expanding universe was smaller and denser, it was
much hotter. Starting from the big bang as the universe expanded,
roughly one hundred thousand years later the temperature cooled
such that its energy was enough for electrons and protons to bind,
forming hydrogen. During that moment the last photons would be
liberated into space, leaving a thermal afterglow at the energy
associated with the binding energy of hydrogen. This is the last fossil
light of the early universe that would persist to propagate throughout
the universe with a temperature of three thousand degrees Kelvin.
But as the universe expanded by a factor of one thousand since the
CMB epoch, we should expect to find this light radiation in every
direction at this temperature. Physicists were on the hunt to find this
afterglow, which was coined the cosmic microwave background
radiation (CMB). Finding it would be a smoking gun confirmation of
the expanding universe. The Nobel Prize–winning discovery was
made in 1967 by Arno Penzias and Robert Woodrow Wilson, two
Bell Lab scientists who first thought that the signal was a
contamination from pigeon dung. Down the street at Princeton,
physicists such as Robert Dicke and James Peebles were looking for
the afterglow radiation from the big bang, but with no success.
FIGURE 11: This is a Penrose space-time diagram of an expanding
universe (top) and a collapsing one. The jagged lines represent the big
bang (crunch) singularity, and the solid diagonal lines represent the
horizon.

When we look in opposite directions in the night sky, we see that


the CMB photons each took 13.8 billion years to travel to us. Those
antipodal photons would take over twenty-six billion years to reach
each other. Because the photons travel at the maximum allowed
speed, we can know the largest distance possible the photons could
have covered, the horizon. Since those photons all existed at the
time the CMB existed, we can ask if they had enough time to speak
to each other at some earlier time. The time between the big bang
and when the CMB existed is three hundred thousand years; we find
that the photons were never able to be in contact with each other. In
other words, they are outside each other’s causal horizon. How is it
that the photons that cannot communicate with each other attain the
same temperature? Some unknown physics that seems to break the
speed-of-light barrier between the big bang and the CMB, which is
about one second in the expansion history, had to take place so that
the photons had the same properties. Otherwise, there is a bizarre
coincidence that gives these photons the same properties that
appear to occur faster than the speed of light. However, this would
violate Einstein’s principle of special relativity—the speed of light is
finite and universal.
But it’s not only the CMB photons that seem to have a nonlocal
origin. After all, if the universe were completely filled with only
photons and electrons, it would be a boring place. How did all the
stars, galaxies, and clusters of galaxies come about? In 1989 a more
precise measurement of the CMB was made and tiny ripples in
photons were found. An analysis of the distribution of these ripples
showed that they matched the superstructure of galaxies in the
universe today—those tiny ripples eventually grew into the stars and
galaxies that occupy our current universe. Here is how it happens.
FIGURE 12: A light cone connecting Earth (at top of image) to the big
bang (at bottom). Three other light cones specify the histories of three
causally unrelated events in the universe’s past.

The physics of the CMB plasma is like a seething hot ocean of a


fluid that’s made up of electrons and photons. This cosmic ocean is
dancing with waves that have a special pattern called scale
invariance. This scale-invariant pattern has a property that the size
of waves of all different frequencies is the same—like zooming in on
the plasma and seeing the amplitude of smaller waves look the
same as larger ones. A random chaotic fluid does not have this
pattern, and some special conditions need to set up an undulating
plasma with this type of scale-invariant pattern of waves. But what is
the purpose of these scale-independent waves? Let’s focus on one
such wave.
Plasma waves are nothing but pressure waves and are plentiful in
nature. Familiar sound waves in the hollow column of a flute is a
pressure wave. When a pressure wave is at its peak the mass
density and pressure of the wave is maximized. In a gravitating
medium like space, these regions of high density will gravitate more
than regions of lower density. While these waves in the CMB are
oscillating from regions of high to low pressure, space is expanding
and cooling the average temperature of the plasma. Eventually the
electrons get captured to form hydrogen atoms and get
gravitationally attracted to the waves of high density. The hydrogen
atoms start to cluster together and become candidates to form the
first stars. The equations that govern the oscillations and infall of the
hydrogen are given by considering a special form of general relativity
called the perturbed Einstein equations. Within a few months of me
writing this, my colleague Jim Peebles won the Nobel Prize in
physics for solving these equations and illuminating the correct
physics that led to the formation of the first structures in the universe
from this nearly scale-invariant spectrum of the CMB’s waves.

FIGURE 13: To the left, during the recombination, epoch electrons and
photons exist in thermal equilibrium. They have the “fluidlike” properties
of a hot plasma. To the right, after recombination, the universe cools and
electrons become bound to the protons, forming hydrogen, liberating the
CMB photons.

The details of how this happens make a more complicated, but


straightforward story. It turns out for Peebles’s equations to correctly
describe large-scale structure of the universe, the CMB also will
have to have extra gravitational pull from overdensities of a new
component of matter, namely dark matter.
Unlike the plasma waves, dark matter is required to have zero
pressure. If this dark matter overdensity has no pressure, then its
oscillation will come to a halt almost immediately. This is because
pressure waves, like springs, resist compression and push back
against compression with more force the more they are compressed.
These two opposed forces of compression and resistance create
periodic oscillations. On the other hand, when the dark matter settles
down into various regions in the sky, the nearby hydrogen will
experience even more attraction toward the dark matter overdensity,
enhancing its tendency to form structures, such as stars and
protogalaxies.
There’s a major problem with this picture, however. The space-
time structure of the expanding big bang universe does not allow
enough time to generate the pattern of scale-invariant waves. So,
something special had to happen before the CMB epoch that set up
all these waves, like a cosmic conductor telling all waves to oscillate
at the same time and loudness. For that to be true, the expanding
universe and the observed properties of the CMB—it even seems as
if our existence—arise from a mysterious nonlocal phenomenon in
the early universe, what Einstein called spooky action at a distance.
Physics acting beyond the horizon and sensitive to a big bang
singularity seems like an inevitable affair to explain our current
existence, and this will require us to consider the possibility of nature
exploiting some quantum magic.
The magic of dark matter isn’t just that it provides the missing
gravity necessary to hold the universe we see together. Many of us
take for granted that next year will come, because we assume that
our solar system’s orbit is stable—without dark matter it is not.
8

A DARK CONDUCTOR OF
QUANTUM GALAXIES

The energy in the undulations that grew under gravitational collapse


some fourteen billion years ago needed an invisible form of matter to
efficiently form the billions of galaxies including our own Milky Way.
We can measure the effects of dark matter in different ways. The
rotation of stars in galaxies is the most direct. Today we see the
striking presence of dark matter in all galaxies, which is inferred from
an anomalous rotation.1 How is this? When we look at gravity at
distances comparable to our galaxy, it fails to account for the motion
of our sun and solar system around the Milky Way. The speed of a
star like our sun around our galaxy is proportional to the amount of
mass it contains. From this, we can infer from the sun’s velocity that
about 85 percent of the mass is missing or invisible/dark. In other
words, our sun is moving so fast that if there weren’t some hidden
form of matter providing the necessary gravitational pull, it would
fling off from the Milky Way into oblivion. We have even found that
every galaxy is made up mostly of dark matter.
A second way we can measure the effects of dark matter is
through a phenomenon called gravitational lensing. Similar to the
bent lens of a magnifying glass, a gravitational lens is a region of
warped space that causes light that traverses it to get distorted.
Imagine that there is a massive blob of dark matter in front of a
galaxy. Although we will be able to see a visible galaxy that lurks
behind the dark blob, the image of the visible galaxy will appear to
be distorted in a definite way according to general relativity. We can
infer the mass of the dark blob from the lensed image of the visible
galaxy.
So, we now know from observations that dark matter exists in
individual rotating galaxies and clusters of galaxies that form a
cosmic web of interconnected galaxies spanning cosmic distances. I
fooled a neuroscientist who thought that a picture of this structure
looked like how the brain is wired with neurons. There is invisible
dark stuff that pervades the universe, and aside from being the
cosmic glue that keeps stars like our sun in orbit, invisible things
have a way of being taken for granted. So why are physicists and
astronomers so interested in dark matter? We believe that unveiling
dark matter is a puzzle piece that will help us understand the
fundamental nature of our physical world. And we suspect that once
we crack the dark matter code, we will come to know something
unanticipated. After all, did Einstein ever imagine that the quantum
nature of the photon would lead to solar cells? Ever since its
discovery, we have been cooking up theories to account for dark
matter, and despite our efforts, we have not been able to identify the
one true model behind its mystery. We are not short of imagination,
as there are hundreds of candidates for the identity of dark matter.
Let’s say that you wanted to construct your dark matter model. There
are some necessary criteria that it needs to satisfy.
Our standard big bang cosmology suggests that dark matter may
have been born in the very early universe at least fourteen billion
years ago along with the visible matter, and based on the
observation of the cosmic microwave background radiation, both
dark and visible matter were distributed across the universe soon
after their creation. After the dark matter is created, it must remain
stable throughout cosmic history, meaning that it cannot annihilate or
decay so that it can cluster to form the scaffolding for the visible
matter to form stars and galaxies. Second, for the dark matter to do
its job to aid gravity in sculpting the cosmic structure, it must be cold,
or equivalently, the dark particles must be motionless, which endows
dark matter with the properties of a pressureless, frictionless fluid.
The opposite of this is a gas of particles bouncing around so fast that
their high momentum would make them hot, which would not allow
structure to form.

FIGURE 14: An artist’s impression of the three-dimensional map of the


network of large-scale distribution of dark matter in the universe.
Apparent in the map are filaments of dark matter that continue to collapse
under gravity to seed galaxies.

As a result, the pressureless dark matter fluid does not exert any
preferential flow in any direction, so it would rather stay put—much
like an invisible cosmic Silly Putty. The dark matter fluid, during the
earliest stages of cosmic structure formation, does undergo localized
undulations, creating distortions of attractive gravitational energy in
the fluid. A correct understanding of the origin of these undulations of
the dark matter is still a matter of research and will be discussed in
the chapter on the physics of the very early universe. The energy in
overdense regions of dark matter generates a gravitational attraction
to enhance its gravitational pull on the nearby visible matter that was
around in the early universe, such as electrons and photons. The
visible matter will collapse around the center of the location of dark
matter density, eventually forming the region for a protogalaxy. The
process I just described is more complicated since it is quantitively
described by a set of coupled differential equations describing how
dark matter sources gravity, and how the visible matter exerts
radiation pressure to create oscillations as it collapses.2 The system
quickly becomes nonlinear and requires numerical physics to
analyze thoroughly. But the takeaway is that once we specify the
dark matter, visible matter, and general relativity, we get the correct
predictions for the imprint of this physics in the observed
temperature fluctuations in cosmic microwave background radiation.
So how do we make progress given the limited things we know
about dark matter and the plethora of theoretical models on the
market? Most models fall into two categories. One is that dark matter
is simply a new particle that’s not found in the pattern of fields in our
standard model. This dark particle/field will just interact very weakly
with the visible matter. Recall that these patterns follow from
symmetries that relate these fields to each other. The other theme
for dark matter is that there is no dark matter, but the force of gravity
changes in the contexts where we see the effects of dark matter. We
call these models modified Newtonian dynamics, or MOND, a
hypothesis created by the physicist Mordehai Milgrom.
The basic idea of MOND is that Newtonian gravity is valid on
solar system scales. Still, as we go into the regime of galaxies,
gravity gets modified to account for the observational fact that all
spiral galaxies rotate at a universal acceleration that is proportional
to the expansion rate of the universe. Galaxies are gravitationally
bound and isolated systems that are decoupled from the expansion
rate of the universe. A typical galaxy is on the order of one thousand
parsecs, and a cosmic expansion occurs at distances around billions
of parsecs. So how could galactic rotation communicate with the
universe’s expansion rate? The jury is still out, but the scientific
community leans toward dark matter being a new particle due to the
observation of the bullet cluster—two galaxies that collided with each
other—leaving in its wake a separation of dark matter and the visible
matter of the two galaxies. This observation is tough (but not proven
to be impossible) to be explained with MOND.3

During most of my college and graduate school years, I was a


competitive 800-meter runner. I trained religiously for three hours five
days a week to meet my goal of finishing the half-mile in one minute
and 51 seconds. Over the years, I would coincidentally see the
number 151 appear when I looked at the time, the number on a
billboard, and many random places.
The number 151 seemed to appear at a frequency greater than
any other number. I took this coincidence as an indication that one
day I was bound to run that time, but twenty years have gone by and
I still haven’t. These days my knees would not even get me through
four hundred meters in two minutes. Had I not given so much
significance to the coincidence of the number 151, I would have
settled my goals on a more realistic 154 for the 800. But aside from
sports and gambling, the seduction of coincidences also creeps up in
physics.
One of the themes in attempting to solve a problem in theoretical
physics appears in coincidences. In the dark matter world, a few
coincidences are floating around. When we try to make progress
using a coincidence, it’s crucial to be mindful of our tastes and
prejudices. One of the most popular coincidences led to the
possibility that dark matter could be detected directly from its
interaction with matter subject to a weak nuclear interaction. The
early universe is a hot thermal soup of particle species. If dark
particles exist and interact, then they will also exist as a thermal
soup as well because the density increases in the past universe.
Most particles, including dark particles, come with their associated
antiparticle that can annihilate into lighter and more stable particles
like electrons and photons. Still, this process is suppressed as the
universe cools and expands, since it is less likely for particles to find
their antiparticles as they run away along with the expanding
universe. There is a beautiful relationship between how the decay of
a heavy particle into a light one will depend on the mass and energy,
and this applies the same way for both dark and visible matter.

FIGURE 15: A schematic representation of the dark matter halo


surrounding the Milky Way, with a supermassive black hole at the center.

What is the lightest mass and interaction strength necessary so


that the decaying soup of heavy, dark particles can give the correct
density of dark particles in the universe today? It turns out that the
mass of the dark particle is related to that of the weak interaction of
visible matter. This is a numerical coincidence and is famously called
the WIMP (weakly interacting massive particle) miracle.
There is another good reason astrophysicists took the WIMP
miracle seriously. One of the great successes of the standard model
of particle physics is the Higgs mechanism, which endows mass to
all the particles known to us. The Higgs field accomplishes this feat
by breaking the symmetry associated with the weak interaction at the
energy scale commensurate with the Higgs particle. A big reason for
building the Large Hadron Collider was to search for new anticipated
particles at the mass scale of the Higgs mass, which would give
insight into the nature of electroweak symmetry breaking and the
physics beyond the standard model.
WIMPs have created excitement because of their prospect for
being directly detectable in particle colliders and other creative forms
of direct detection here on Earth. For example, every once in a while,
a WIMP can travel to Earth from a galactic halo and interact weakly
with an atomic nucleus, which could give a detectable signal
according to the specific type of WIMP particle and its interaction
with the nucleus. My colleague Rick Gaitskell at Brown has been one
of the pioneers of the quest to detect WIMPs and other dark particles
directly. It’s been about forty years, but so far, and with ingenious
and laser-sharp accuracy in detecting the faintest interaction with
WIMPs, there has been no direct detection. These direct-detection
experiments have allowed us to highly constrain and rule out a large
class of WIMP models. Still, a number of models, predicting a variety
of possible particles to fill the role of WIMPs, survive, and the hunt is
still on to see which of those particles might actually exist. There are
some interesting anomalies among the different experiments that
suggest that WIMPs may be the culprit. I remember having lunch not
too long ago with one of the co-inventors of the WIMP miracle
model. She lamented to another dark matter theorist, “It’s been over
thirty years; where are the damn WIMPs?” So, let’s assume that the
dark matter is not a WIMP. Then what are we left with? For one
thing, if it is not a particle, then what else could be in the sky?

One uneventful fall day after teaching my general relativity course, I


noticed an email from my colleague David Spergel, a pioneering
cosmologist, director of the Simons Center for Computational
Astrophysics, and now president of the Simons Foundation. He said
that he would be passing through Providence and would have a free
half-hour to meet. Spergel was one of the first people to notice the
WIMP miracle while he was a young graduate student and later with
renowned cosmologist Katie Freese figured out an ingenious way to
detect WIMPs. We decided to have a quick dinner close to the
Amtrak station in downtown Providence.
I knew that Spergel was a tad frustrated that WIMPs had not been
detected after decades of relentless searches, and I saw this as an
opportunity to pursue viable alternatives with him. Over a course of
Malaysian food, we wasted no time, and I asked him: Aside from the
WIMP miracle, what was the most pressing observational mystery
behind dark matter that needed to be addressed?
He reminded me of another coincidence, an observation that he
was among the first to notice. The observed density of dark matter is
the same magnitude as the visible matter density—and a
fundamental explanation was lacking.4 This coincidence is even
more striking because visible and dark matter don’t seem to interact
with each other.
After going back and forth, we realized that the link between dark
and visible matter could be addressed by their common origin in the
early universe. The realization was fueled by the need to combine
the quantum principle with the gravitational side of dark matter. As
we will soon see, not only does dark matter gravitate but gravity
plays a key role in the origin of dark matter, and this would have
problems without quantum mechanics. A part of this issue was
pursued and addressed by particle physicist David Kaplan in the
1980s.5
Kaplan is known to be one of the most creative particle physicists
in his generation, and his idea called asymmetric dark matter was
the first to connect creation of matter over antimatter to the creation
of dark matter, attempting to explain the coincidence between their
densities in the universe. But, how could it be that dark and visible
matter can have a joint creation without ever “knowing” about each
other? Plus, it is not enough to provide a candidate for dark matter
that solely explains its properties in the current universe. It is useful
to understand baryogenesis itself to answer the potential link behind
dark and visible matter. Even if we were to understand the origin of
the dark matter, David Spergel and I understood that our model was
calling for something that was not a WIMP. It is first useful to
understand how dark matter can be cocreated with the visible matter,
otherwise known as baryogenesis—the origin of baryonic matter,
such as protons and neutrons.
As it stands, there is no consensus among cosmologists about
the creation of matter over antimatter in the universe. But there are
some conditions that most of us agree on, which have to do with the
known symmetries of the vacuum state of the quantum fields that
give rise to all the matter and antimatter. The critical insight into
baryogenesis is that special symmetries that related matter to
antimatter had to be desecrated by new physics that operated in the
very early universe. And as we will see, these symmetries are not
cherished in our universe. In 1967 the Russian physicist Andrei
Sakharov provided the necessary ingredients to accomplish
baryogenesis—these were famously known as the Sakharov
conditions. These conditions fit into the principle of emergence that I
discussed earlier, where we see that the violation of invariance leads
to a variety of complex, emergent, and physical properties.
Baryogenesis transpires to be a most crucial cosmic example of the
emergence of matter in space-time, but exactly how the universe
does it remains a mystery.
To create matter over antimatter in the universe, dark or visible,
Sakharov imagined that three symmetries would have to be
disrupted. The vacuum of the quantum fields in the standard model
has a symmetry that preserves the number of particles. So, if the
universe starts empty, then it will remain barren unless there is a
process that violates the conservation—this is called baryon number
violation, or B. But even if we have a process that violates B, nothing
stops that same process from producing antibaryons. This is seen in
the following example: An electron and positron both democratically
interact with a photon. If enough energy in the photon is imparted
into the electron’s field, then its vibrational energy will enable the
creation of an electron from the vacuum. But, even if you were a
demigod trying to create a certain number of electrons in the
universe this way, you would necessarily create just as many
positrons as well. But if you want to create a universe like ours, there
needs to be some way around this. This is Sakharov’s first condition:
the baryon number cannot be conserved.

FIGURE 16: Representation of CP symmetry violation in a pair of


electrons.
There is another symmetry between the electron field and the
photon that also has to be violated. Called CP, where C and P stand
for charge and parity inversion, this symmetry is similar to a mirror
reflection. When you look in the mirror, there is a spatial reflection
that inverts left and right—your left hand looks like your right hand
under a mirror reflection. We can imagine CP inversion as a kind of
mirror that reflects the charge and orientation of the electron. A
spinning particle under a CP inversion sees itself looking like its
antiparticle, the positron with its spin reversed. So, if we wanted to
create matter over antimatter this CP symmetry would have to be
violated. This is Sakharov’s second condition.
The third condition is that both charge asymmetry and particle
creation would have to occur out of equilibrium, meaning that
creating a particle would have to happen such that the particle would
not encounter its antiparticle as they would demolish each other. But
when in the history of the universe did these conditions occur for
creating matter over antimatter? Attempting to understand the origin
of both visible and dark matter points us to an earlier epoch in the
early universe called cosmic inflation that had the correct physical
ingredients to realize Sakharov’s conditions and a universe that is
void of matter for all intents and purposes. We will get to how
inflation works its magic creating “something” from “nothing,” but let
us return to our discussion about dark matter genesis.
David Spergel and I teamed up with my postdoc and fellow
cosmologist Evan McDonough to develop a new theory of dark
matter that realizes the coincidence between dark and visible matter
baryogenesis while being agnostic about the underlying field theory.
We let a quantum consistency condition, known as anomalies,
dictate what the nature of the dark matter would be and discovered
two pleasant surprises.6 First, we discovered that the dark matter
originated from a mirror world of the quarks in our visible world—
dark quarks. After they are produced these quarks are close to zero
temperature and condense, similar to how helium atoms condense
to form a quantum fluid known as a superfluid. The dark quantum
superfluid further clumps up under gravitational collapse to make
galactic blobs that will differ from ordinary WIMP dark matter. We
were surprised about this quantum mechanical result that our
calculations revealed, but we were not the first to resort to dark
matter as a galactic quantum phenomenon.
Two decades ago, my colleague and friend Wayne Hu, who did
pioneering work on relating the CMB fluctuations to fundamental
physics, made a quantum leap in our imagination of dark matter.
This realization arose from paying attention to a dark secret about
dark halos and how they act as a “conductor” for the motion of visible
matter in galaxies. There is an entire research enterprise that uses
supercomputers to simulate the formation and evolution of galaxies
spanning billions of years. These so-called large N-body simulations
computationally study the cosmic history of dark matter particles that
lead to their formation into dark halos. Time and time again with
unprecedented accuracy, large N-body simulations reveal that the
distribution of the dark matter in these halos has a cuspyness in its
concentration around the center of the galaxy, in disagreement with
observations of real halos, whose dark matter distribution is smooth.
Hu took this discrepancy seriously and asked what it would take to
prevent the dark matter from clumping in the center of galaxies. And
his answer was both simple and struck a conceptual resonance with
how Newton connected the falling of the apple to the motion of the
moon. Hu saw that the same wavelike property that stabilized the
electrons’ orbit in the atom can prevent dark matter from forming
cusps, if the dark matter had a quantum wavelength on the scale of
a galaxy!
Recall that the de Broglie hypothesis says that a massive
quantum particle should have a wavelength inversely proportional to
its mass. This means an extremely light particle will have a long
quantum wavelength. The reason we don’t see the wavelike
properties of particles like electrons and protons is because their
masses are large enough that their de Broglie wavelengths are
microscopic. What Hu argued is that the dark matter can have a
quantum wavelength on the scale of a galactic distance provided
that it is light enough. And like a galactic electron the light dark
matter will not collapse to the center of the galaxy.
Unlike the plethora of possible WIMP particles, so far there are
only two potential candidates for superfluid dark matter, a particle
that we’ll discuss later called the axion and the dark quark
condensate. We are still trying to find new, clever ways of unveiling
the true identity of dark matter, assuming that it is a form of matter
that doesn’t interact with matter that we see. In the case of superfluid
dark matter, we can exploit its quantum properties as a way of
distinguishing its properties revealed in the sky. And one unique
property of superfluids is that they produce quantum excitations
called vortices, which have a mathematical pattern similar to how
clouds spiral to make the eye of a storm. If galaxies have vortices in
them then they would affect the gravitational lens effect and lead to
an anomalous observable signature. Currently, research groups,
including my own, are using machine learning to hunt for
substructure such as vortices within dark halos as a smoking gun
signature of the quantum nature of dark matter.
9

COSMIC VIRTUAL REALITY

The ancient idea that the universe was created from a state of
emptiness is captured poetically in the book of Genesis: “And the
world was without form and void, and darkness was on the face of
the deep. And God said, Let there be light, and there was light.”
Similar sentiments were expressed earlier in ancient Babylonian,
Sumerian, West African, and Maori creation philosophies. Our
precision physics of quantum field theory and general relativity
combine to give a picture of the early universe that resonates with
these creation stories—the idea that everything emerged from the
void. The notion of emptiness, of blackness, according to modern
physics, is not how we imagine it to be based on our direct
experience of empty space. In every region of empty space,
quantum fields are seething with activity so rapid that our ordinary-
day perceptions are blind to it. But empty space is much more
interesting that our psychological and cultural projections make it to
be. The modern picture of the early universe is that the ignition of the
big bang emerged from a previous primordial state in which the
world was devoid of matter in a vacuum state. My goal here is not to
turn our discussion into one of a perennial matter but to motivate the
underlying physics that presently attempts to explain the creation of
the physical universe from an empty state during the big bang
epoch, and discuss some current mysteries cosmologists face.
To date, the most accepted and experimentally compelling
paradigm of the early universe is cosmic inflation. The critical
message of inflation is that every bit of structure in the universe,
including planets and living things, emerged from quantum
fluctuations from a vacuum state that contains the inflaton field’s
potential energy—the inflaton is called that because it drives inflation
—and nothing else. Nevertheless, what physicists call virtual
particles can emerge spontaneously from the vacuum, and these are
important to explaining how both matter and large-scale cosmic
structure emerged from nothing. And although there are alternatives
to cosmic inflation, such as the big bounce and cyclic cosmologies,
which attempt to alleviate the problems that plague inflation (we will
discuss some of these alternatives in the next chapter), none of them
are free from the influence of virtual particles. No matter which way
we slice it or which model we prefer, the idea that all matter and
structure in the universe came from virtual quantum processes is
inescapable, counterintuitive, and requires a more in-depth
examination. After all, nature has to figure out how to make virtual
particles real particles. And indeed cosmic inflation has a clever trick
up its sleeve to make this happen.
In quantum field theory, virtual processes are ubiquitous and have
been measured in the lab. A consequence of these virtual processes
are virtual particles, which are particles that quickly materialize and
disappear into the vacuum before having any material
consequences. The vacuum is not empty but seething with rapid
interactions of quantum fields and particles. Remember that
quantum fields comprise oscillator modes that vibrate in the vacuum
state. These field oscillators also interact with other field modes. A
virtual process occurs when field modes spontaneously activate to
create a particle, and like a fish leaping out of the sea the virtual
particle quickly returns back to the vacuum state. An example of this
is when an electron and positron get pair produced from field
oscillations and almost immediately annihilate into the vacuum state.
Therefore under normal circumstances, these virtual particles never
materialize into real ones.
It is the uncertainty principle that is at the heart of virtual
processes, and it is remarkably dictated by the following relation,
which says that the uncertainty in the energy of a virtual process is
inversely proportional to the uncertainty in time that the process
occurs. According to the uncertainty principle, extremely microscopic
time scales correspond to very high energies, enough to pair
produce matter and antimatter from the vacuum state. For all
practical purposes, the vacuum is continually creating and destroying
particles at a rate so fast that we do not “see” them because they are
too short-lived.
Using Einstein’s relation that equates energy with mass, we can
find the condition to create two particles of a given mass out of a
field that carries the necessary amount of energy. As an example, if
we combine Einstein’s energy-mass relation, stated above, along
with the uncertainty principle, we find that it is possible to create
electrons and positrons during a time interval that is roughly a
trillionth of a trillionth of a second! During that time the particle and
antiparticle pop into and out of existence (because they soon
annihilate). That’s why we never see virtual processes in everyday
situations. As we will now see, according to inflationary theory, all the
matter that exists around us, including ourselves, were originally
virtual particles. How is this possible?
In everyday situations where the curvature of space-time is nearly
flat, virtual quantum fluctuations are incredibly short-lived and remain
quantum—they never take on all the characteristics of a classical
particle. But inflation produces these virtual particles, and somehow
they do become classical. Recall that in quantum mechanics, the
collapse of the wave function requires a mechanism to solve the
measurement problem. The Copenhagen interpretation postulates
that a classical measuring device or observer collapses the wave
function. However, there are no known observers or classical
apparatus during or soon after inflation that we can postulate to
collapse the wave function of these virtual states. Therefore, inflation
has a cosmological measurement problem.
FIGURE 17: A Feynman diagram of a virtual process. The wavy line
represents an incoming photon, which emits a virtual particle,
represented by the closed loop.

However, due to the dynamic nature of the expanding space-time


during inflation, virtual processes play a central role in becoming the
very seeds of structure in the universe. Let’s take a deeper dive into
the virtual processes during inflation. One interesting phenomenon is
the exponential expansion of space-time feedback into the dynamics
of the quantum vacuum fluctuations, causing them to reduce their
uncertainty. The reduction in uncertainty transmutes the oscillations
to be a coherent state. This might sound fanciful, but in fact, these
states are identical to when photons collectively become a coherent
laser beam. These coherent states undergo a phenomenon called
quantum decoherence, which washes out all the quantum phases in
the wave function, rendering the primordial quantum fluctuations into
classical seeds for cosmic structure. The key to virtual processes
during inflation is that the time during which inflation happens is
shorter than the time scale, dictated by the uncertainty principle,
necessary to produce the observed particles around us. Inflation can
last as quick as a millionth of a billionth of a billionth of a second.
During this short time, the inflaton dumps its potential energy into the
production of virtual quantum particles. These virtual particles live
long enough and simultaneously get stretched to transition from
virtual particles to real classical particles such as electrons, quarks,
photons, and others produced in the big bang.

FIGURE 18: Schematic representation of inflation in the universe.

Cosmic inflation was invented by my mentor Alan Guth to solve a


handful of problems that plagued the standard hot big bang
paradigm, especially the horizon problem that we discussed earlier.
Inflation’s magic stems from a short burst of exponential expansion
of the universe, transforming a microscopic universe into the
macroscopic one we inhabit. Soon after, the inflationary theory was
improved by Andrei Linde, Andy Albrecht, and Paul Steinhardt to
address some technical problems in Guth’s original proposal. The
potential energy for inflation is ignited during the Planck epoch,
which is the earliest stage of the big bang, during which quantum
effects are expected to reign supreme due to the universe’s
microscopic size and high energy scales. A large amount of potential
energy can transform into kinetic energy of the gravitational field,
creating an exponentially fast expansion rate. The consensus from
theorists is that this energy is contained in a new spin-zero quantum
field called the inflaton. On the largest scales, this field has to be
smooth and homogenous to create the observed homogeneity seen
in the night sky: like a well-shaken bottle of milk, the sky looks pretty
much the same whichever direction you choose to look. During the
rapid expansion, a few magical things simultaneously happen. First,
all present quantum field oscillations get created from the vacuum
and stretched by the expansion. This happens because the part of
the gravitational field that expands, known as the scale factor,
directly couples to the wavelengths of quantum vibrations, acting like
a volume amplifier in a cosmic stereo. As the scale factor increases,
wavelengths likewise get stretched.
This basic picture of cosmic inflation is compelling and makes the
correct prediction of the observed properties that we measure in the
fluctuations of the cosmic microwave background radiation. Indeed,
the so-called empty space in the early universe, if it is dominated
with vacuum energy, comes alive through the engine of inflation, to
convert this potential energy into the observed particles and radiation
we see. The magic behind inflation is its effect on general relativity to
create an exponential expansion of space itself. All that existed in
this infant universe were quantum fluctuations riding along with this
inflationary era. And it is their particle excitations that get created out
of the vacuum of space-time that becomes all the observed matter
around us, including us. However, there remain some foundational
problems of inflation that need to be addressed.
FIGURE 19: The wiggly lines represent the quantum density fluctuations
sourced by the epoch of cosmic inflation. At later times these density
fluctuations grew, with the help of dark matter, to form galaxies.

The quantum birth of matter and antimatter during inflation is a


very promising avenue for generating the observed asymmetry
between matter and antimatter that characterizes our universe today
—that is, for solving the problem of baryogenesis. Inflation does the
job of producing matter, but it also produces antimatter. Unless
matter over antimatter can be produced either during inflation or
after, we are left with a world of equal matter and antimatter, and
that’s not our world. If there were a way to bias the production of
matter, then we would explain a big part of why we don’t see
galaxies, planets, and stars made of antimatter. There have been
heroic attempts to account for baryogenesis in the early universe,
and most models assumed that the necessary ingredients occurred
after inflation, and this made good sense. The rapid expansion of
inflation expands the volume of space so much that an initially large
number of particles would dilute away. But aside from this issue, two
conditions spelled out by Sakharov are naturally satisfied during
inflation—particle production and the out-of-equilibrium condition.
The key lesson behind inflation is that our classical world of
galaxies, stars, and planets emerged from a quantum origin.
However, the baryon asymmetry was established either during
inflation or by new physics after inflation. In 2004 my collaborators
Michael Peskin and Mohammed Sheikh-Jabbari and I developed a
theory that shows how inflation can solve the problem of
baryogenesis. And this idea, many years later, also provided a way
to generate a specific breed of dark matter. The key to this
mechanism is to realize that the very agent that ignites inflation, the
inflaton field, can interact or couple with matter and antimatter in an
asymmetric manner that violates CP and baryon number in one fell
swoop. This is elegant because instead of struggling to satisfy the
three Sakharov conditions independently, the inflaton meets them all
in one shot.
That doesn’t solve all the challenges facing inflation, however.
Inflation assumes that quantum fields, including the inflaton field
itself, are subject both to the laws of quantum mechanics and to
gravity. Gravity is mostly classical. However, for inflation to do its
magic, some aspects of gravity have to be quantum mechanical as
well. The seeds of structure, including the virtual processes, are
genuinely born when they interact with quantum undulations of the
gravitational field. These undulations are small compared to the part
of the gravitational field that undergoes homogeneous growth in
three spatial dimensions. But we are free to ask, why not quantize all
gravity? Nothing stops the tiny quantum fluctuations of the
gravitational field from growing the closer we get to the time that
inflation begins. Why should we fully quantize the matter fields and
not fully quantize gravity?
The quantum fluctuations of inflation also involve quantum
fluctuations of the gravitational field. And as we go back to earlier
times when inflation itself was ignited, these quantum gravitational
fluctuations approach a big bang singularity hinting at a need for a
full treatment of quantum mechanics and gravity, a theory of
quantum gravity. We will see that quantum gravity and inflation
forces some sharp questions about assumptions we make about
physical reality. But to get to the question of gravity, let’s consider
something like antigravity first.
10

EMBRACING INSTABILITIES

A global virus pandemic brings the world to its knees. The stock
market suddenly drops, inciting fear from investors. A star collapses
to form a black hole. These events all have one thing in common.
They are all instabilities that correspond to a catastrophic growth in
some quantity that leads to an unwanted outcome. In physics,
sometimes the bad outcome of an instability threatens to obliterate
the validity of the theory itself. The quantum revolution was born in
part as a result of taming instabilities in atomic systems, such as the
ultraviolet catastrophe and the instability of classical atomic orbits. In
recent times, billion-dollar particle accelerators were built to look for
supersymmetric particles that function to tame an instability that
would lead to a catastrophic growth in the masses of all matter. No
diet would help us for such an instability. Many physicists were
confident that this pattern of fixing instabilities would lead to
experimental confirmation of supersymmetry, but it didn’t happen.
Are all instabilities tragic, or are some useful for our universe’s
functioning in hidden ways that could lead to new directions in our
understanding? What are instabilities, especially of the quantum
type, trying to tell us about the new physics?
My colleague João Magueijo at Imperial College developed a
theory of the early universe that solved the infamous horizon
problem and flatness problem in cosmology. In ordinary relativistic
cosmology, the horizon problem describes the fact that the edges of
the universe seem to have been in communication with each other,
despite being too far apart for light to have traveled from one far-
flung region to another in the time the universe has existed. The
flatness problem has to do with why there’s just the right amount of
energy in the universe so that its geometry is Euclidean and flat. In
Magueijo’s theory, he posited that in the earliest stages of the
universe, the speed of light could vary in time, at odds with Einstein’s
postulate of the constancy of light. Magueijo presented a
mathematical model of his theory in a publication that was later
contested by some theorists as being a sick theory, because it
seemed to give rise to instabilities if the theory were to be quantized
—to some aficionados, it was not even wrong. I was there for the
showdown, and Magueijo, being a black belt, enjoys these physics
sparring sessions. We were sitting at a seminar and the invited
speaker made reference to Magueijo’s theory and claimed that it was
sick because it had an instability. Magueijo responded, “I want the
damn instability. After all you are an instability!”
What Magueijo meant was that the chain of cosmic events
leading to life were ignited by instabilities. These tiny perturbations in
the CMB that grew to form the first structures is an instability in the
equations that govern those cosmic structures. These good
instabilities get regulated and stabilized when the system becomes
nonlinear and highly interdependent with the gravitating
environment. Today physicists are wrestling with strange instabilities
that continue to cause the physics community headaches. We will
place our focus on an instability that is found in empty space, a
vacuum instability that communicates with everything otherwise
known as the cosmological constant or dark energy problem.
Recall that the discovery of quantum mechanics was ignited by a
handful of instabilities found in classical physics. Our very existence
owes to the stability of the atomic structure of hydrogen and oxygen
in water molecules, but classical physics predicts that all electrons
should spiral into protons, making the classical atom unstable. Every
time the electron orbits around the proton it radiates away
electromagnetic energy, which reduces its distance to the proton.
Eventually it spirals into the proton, rendering the atomic system
unstable. By quantizing the orbits of the electron by associating each
orbital distance with a wave, the electron, like the lowest vibration on
a guitar string, would have a lowest orbit allowed.
Another type of instability is when a system’s energy continues to
grow without bound. This is similar to falling down an infinite hill—
your kinetic energy would continue to increase until it reaches
infinity. And this contradicts relativity, which says nothing can go
faster than light. Ironically, while quantum mechanics was invented
to tame classical instabilities, it was later discovered that even
quantum systems can have instabilities. For instance, when we
consider the quantum effects of electrons, what we perceive as
empty space in front of us turns out to contain a form of energy due
to the activity vibrating quantum fields. This contribution to energy
(often called vacuum energy) is very large and should dominate the
energy of the universe, causing this expansion to accelerate so fast
that galaxies would not have a chance to form. This is known as the
cosmological constant or dark energy problem, and as of today there
remains no solution. Let’s understand this at a deeper level.
As we explored previously, the fundamental substance behind all
matter and force carriers known to us are quantum fields, and they
all generate vacuum energy. As we saw earlier, one consequence of
quantum fields is that empty space is not empty; it is filled with
vibrations of quantum fields that, in the universe’s past, built all
substance from stars and planets to life itself. Thus, empty space is
occupied with these vibrating, interacting quantum fields. Recall that
a quantum field has properties like a large collection of oscillators.
Because of the uncertainty principle, a quantum field oscillator can
never be at rest.1 For such a system the rate of vibration is
proportional to the field’s energy and can emerge as a particle from
empty space like a water droplet that jumps on the surface of the
ocean.
The baryogenesis problem makes clear that just as every particle
is created, so too is its antiparticle created, but only for a short
moment of time, beyond our ability to see with our naked eyes or our
most advanced high-speed shutter cameras. This spontaneously
created pair of particle and antiparticle quickly meet and annihilate
each other, back into empty space. What’s the use of that? This
process is called a vacuum bubble, and it generates energy. Gravity
likes energy. There are other similar effects that likewise generate
vacuum energy. In every chunk of empty space, the vacuum energy
can be envisaged as completely transparent quantum fluid that
differs from ordinary fluid. If you try to compress an ordinary fluid like
water, it will resist compression and push back on you; this is
positive pressure. The vacuum fluid has negative pressure and does
the opposite. Try to compress it, and it will expand outward. The
effects of quantum vacuum energy in empty space have been
confirmed in the lab and they are tiny. But the predictions of our
standard model predict a much larger amount of vacuum energy.
And all forms of energy according to general relativity, including
vacuum energy, will warp space-time. So, how exactly does vacuum
energy affect curved space-time?
If there is one key lesson to take away from Einstein’s theory of
general relativity it is that space-time is also a field. The matter fields
and their vacuum energy are tethered to the gravitational field,
causing it to warp. The vacuums’ negative pressure stretches space
by accelerating observers away from each other, resulting in a
repulsive force. For instance, if there was a lot of vacuum energy
near the earth and the moon, space would stretch rapidly outward
and they would fly away from each other. Now comes the punch line.
In 1998, groups led by astronomers Saul Perlmutter, Adam Riess,
and Brian Schmidt observed with powerful telescopes how distant,
exploding stars recede away, and showed that the expansion of the
universe is accelerating. The only way this can happen is if the
universe were filled with vacuum energy. But there is a major
problem: according to precision calculations in quantum field theory
and general relativity, the expected and observed vacuum is 120
orders of magnitude smaller than what our standard model predicts.
You may wonder why there is all the fuss about this discrepancy
between our most precise theoretical characterization of modern
physics and our observations. For one, it says that something is
wrong with either general relativity or quantum field theory, or both,
in describing the quantum effects of fields on space-time.
Alternatively, there is some unknown reason why the vacuum energy
is either hiding itself or tamed by some new, mysterious, unknown
physical force—yet to be discovered.

FIGURE 20: Systems can be stable if the potential energy is bounded


from below, as seen to the left. The middle metastable graph shows a
bounded potential locally, but unbounded to the left. So, this particle can
be both stable and unstable depending on its energetics. An unstable
system has a potential that is unbounded in all directions.

The form of vacuum energy that generates the universe’s


acceleration is popularly known as dark energy. Our understanding
of precision physics predicts an instability in the production of dark
energy. In other words, there should be too much dark energy and
the universe should have accelerated much too fast earlier on to
even form stars, galaxies, and us.

After a research stint in London, I headed out to the hills of Silicon


Valley to continue my research at the Stanford Linear Accelerator
Center (SLAC), known as one of the meccas of particle physics. My
new colleagues at SLAC and Stanford were some of the pioneers in
string theory, a theory that realizes Einstein’s dream to unify all four
forces. The idea of string theory is simple and elegant; the
articulation of the theory requires some of the most advanced
mathematics known to us. My new colleagues had worked for
decades along with a global community of string theorists to develop
this elegant ten-dimensional, yet abstract framework. Young
physicists like myself were handed the baton in this quest. String
theory provided a rich framework to address problems that share
residence in particle physics and cosmology. This is mainly because
string theory naturally links the extra dimensions with gravity and
hidden symmetry patterns found in the standard model of particle
physics.
There are strong reasons to believe that string theory could very
well be the theory that underlies our reality. As we will explore in an
upcoming chapter, string theory unites quantum mechanics with
general relativity by taming disastrous ultraviolet (short wavelength)
quantum instabilities.2
Once cosmologists measured the amount of dark energy (also
known as vacuum energy or the cosmological constant) deduced
from the acceleration of space-time, we had to reckon this fact with
our precision theories. We know that the standard quantum field
theory interacting with general relativity was not enough. This was an
opportunity for more fundamental theories such as string theory,
which had ingredients of taming certain short scale instabilities, to
come to the rescue. We saw this as some experimental guidance for
our theories, places where the theories could possibly make contact
with the real world and tell us why the universe was accelerating at
this epoch in cosmic history. However, a major set of challenges
presented itself. String theory is blessed with symmetries that tame
infinities, but the symmetries themselves did not straightforwardly
allow for a positive cosmological constant, let alone the tiny value
that exists in our universe. I approached the head of our group and
asked him for some guidance in how I and the other postdocs might
navigate through the difficulties and confusion. His advice was to find
an example of a solution from string theory that admits any positive
cosmological constant. We were all on a hunt to find this solution.
Within a year my colleagues Shamit Kachru, Renata Kallosh, Andrei
Linde, and Sandip Trivedi found a remarkable pathway and solution
of realizing a cosmological constant in string theory. And this had to
do with finding a special space-time solution in string theory called
de Sitter space.
Solutions of general relativity that admit a positive cosmological
constant were known since the inception of the theory. One solution
was discovered by Willem de Sitter, and as expected, it gave
accelerating space-times; the larger the cosmological constant, the
more the space accelerates, and by an exponential amount. During
the early stages of the universe’s expansion, there was a fine
balance between the repulsive expansion of space and the attractive
gravitational pull of infalling matter to form the first stars and
galaxies. The Nobel laureate Steven Weinberg realized that if there
was too much vacuum energy in the universe, the repulsive force
would overshadow the attraction necessary to form structures and
there would not be any structure, hence there would be no life in the
universe; there would be no universe as we know it. The universe
seems to have dialed in just the right amount of vacuum energy for
structures and life as we know it to exist. So it is up to our theories to
understand not only why vacuum energy is positive but also why the
exact amount needed for the existence of the universe is observed—
we call this the “why now?” or coincidence problem. Unfortunately,
as we’ve seen, our current understanding of physics produces far
too much vacuum energy. So, the real question is: What happened
to all the vacuum energy that we expect to exist? Or let us assume
that all this vacuum energy does exist, so then why does gravity not
respond to vacuum energy? This remains a mystery.
During my time at Stanford, young string theorists like me were
seeking out pathways in the jungle of ten-dimensional calculations to
find solutions that have a small and positive cosmological constant.
Then, in a groundbreaking paper, renowned theorists and friends
Raphael Bousso and Joe Polchinski argued that the search to find
such solutions could be futile. I’ll come back to their work later in the
book; for now it’s enough to say that it essentially argued that the
universe was just one of many, which have a vast range of possible
values for the cosmological constant, without any first principles to
force its value. By reasoning that has come to be known as the
anthropic principle, they argued further that we necessarily live in a
universe with a cosmological constant that is capable of supporting
life like us, and so we shouldn’t seek any deeper explanation for it. A
big debate transpired in the cosmology and string theory community
as to whether the anthropic principle was scientific. I decided to take
another direction, which would risk further shunning from my
colleagues. This new direction would mean that I would engage in
conversations with the outsiders from my club and even import their
ideas into a possible resolution of the cosmological constant
problem. I was guided by the motto: “Let the nature of the problem
dictate the tools you should resort to,” even if it meant borrowing
from the outsiders or risk becoming one yourself. Active deviance
was on the horizon.
That I even wanted to work on the cosmological constant problem
was already deviant behavior. In the aftermath of that paper,
postdocs were warned to not work on it. Some of my advisers
suggested that we postpone working on the cosmological constant
until we got tenure. It was a Medusa that defeated the morale of
many theorists who attempted to decipher one of nature’s biggest
puzzles. But I was haunted by the beauty of the cosmological
constant problem and was fine with joining the ranks of those that it
defeated. Plus, it was my last year on the job market for a faculty job,
and I did not have high expectations of getting a permanent position,
so I did not feel like I had much to lose.
If there was any chance to tread forward, I would have to find a
new direction that was not thought of before. One day while having
coffee on top of a hill in Nob Hill, I saw a similarity between the
cosmological constant problem and a problem that haunted particle
physics, the strong CP problem that came up in baryogenesis and
the origin of our matter-filled universe. In the strong nuclear
interaction, described by a theory otherwise known as quantum
chromodynamics (QCD), the gluon is the particle that mediates the
interaction between quarks that bind to neutrons and protons.
Classically the neutron is electrically neutral, but quantum effects
induced by QCD create a very large amount of net electric charge in
the neutron that would destroy the stability of atoms. This would be
catastrophic, given that we’re made up of neutrons and protons. Like
the cosmological constant in general relativity controlling whether the
universe expands, is stable, or contracts, there is a parameter in
QCD, called the theta parameter, that controls the amount that the
neutron deviates from electrical neutrality. Experimentally the theta
parameter was measured to be on the order of one billionth. So, the
strong CP problem is relegated to a question as to why the theta
parameter was so close to zero, a similar predicament as the
cosmological constant.
In the late seventies, Helen Quinn and Roberto Peccei found an
elegant solution to the problem. They realized that the strong
interactions could have a hidden new symmetry. Think of this
symmetry like a particle rotating around a frictionless surface in a
perfect circular orbit. If all forces were absent, every point of the
ball’s orbit would have the same energy. So, the energy has a
symmetry such that any rotation of the ball leaves the energy
invariant. Now if we add gravity to the problem and slightly tilt the
circle, then the gravitational force would break the symmetry, simply
because there is potential energy in the gravitational force. Similarly,
Peccei and Quinn showed that a quantum effect—the emergence of
a new field called the axion—breaks CP symmetry, introducing a
potential energy. This axion readjusts itself to the minimum of the
potential and conspires to drive the theta parameter to zero. I
wondered if we could reimagine the cosmological constant to act like
the theta parameter of gravity. Luckily, Helen Quinn was at SLAC. I
told her the idea once over lunch.
We sat at a work desk in her office and Quinn meticulously
walked me through the conceptual and mathematical inner workings
of her solution to the strong CP problem. There is a big difference
between reading a research paper on a physics result and learning it
directly from the author herself. It was only when I saw how Quinn
thought about the strong CP problem and her and Peccei’s
ingenious insight into the solution that I was able to continue to find a
way to implement the idea into gravity and the cosmological constant
problem. Most importantly, while Quinn had high standards for the
creative and technical implementation of the idea, she was
encouraging, and it empowered me to take the next step.
To take the analogy between a problem in QCD and gravity to a
place where I could attempt to do a calculation, it would help if I
could place gravity on similar footing with QCD, and there was one
activity of research that already did that. In the arena of quantum
gravity, string theory is seen as the only game in town, but it isn’t.
There are other attempts to quantizing gravity, even if they do not sit
well with my string theory friends.
One particular approach is loop quantum gravity (LQG), in which
the starting point is to quantize gravity using the same methods as
QCD. This possibility came from Abhay Ashtekar’s ingenious insight
to rewrite general relativity using variables identical to QCD.3 The
idea required me to use the Ashtekar formulation of gravity in the
presence of a cosmological constant. When I spoke to the other
postdocs about loop quantum gravity, many of them dismissed the
theory as “loopy” and suggested that anyone that would work on that
theory does not know physics. But I already felt like an outsider and
pursued loop quantum gravity anyway. Besides, when I pressed
some members of the group to provide a solid critique about loop
quantum gravity rather than just make fun of it, most of them did not
know the theory well enough to tell me why they thought it was
wrong. So, I decided to invite one of the founders of loop quantum
gravity, Lee Smolin, to come to Stanford and SLAC to give some
lectures on the theory. That way at least we could have a more
informed critique of the theory as a group, and I could make
progress on my idea for the cosmological constant. I finally learned
enough about the Ashtekar formulation to get going on the project.
But my invitation to let an outsider come into our club to teach us
something left a bad taste in the mouths of members of my group.
That was just one of the reasons I spent much of my time at a
café called Cup of Joe near the top of Nob Hill performing
calculations on the project. I would frequently drop by the office of
my adviser, Michael Peskin, to discuss my work when I got stuck.
Finally, I finished the project and provided a mechanism to partially
solve the cosmological constant problem in a manner similar to the
Peccei-Quinn mechanism. The key idea is that there is a new axion
that, like a person wobbling on a high wire, can readjust itself to
cancel the cosmological constant. The model still needed more
improvement to address the other aspects of the cosmological
constant problem. However, my results were enough to put the
paper up for publication, and the idea inspired a new way to think
about resolving the disastrous production of dark energy.
The paper, entitled “A Quantum Gravitational Relaxation of the
Cosmological Constant,” was posted on ArXiv.org, a website where
physicists share drafts of their papers with the global physics
community. Days of silence from the community went by. I was not
offended. My office mate, string theorist Amir Kashani-Poor, returned
after giving a seminar at the University of Texas at Austin. He told
me with a look of awe that he had been lunching with the string
theorists, and Steven Weinberg had joined the group. Weinberg
shared the Nobel Prize with Abdus Salam and Sheldon Glashow for
unifying the electromagnetic force with the weak interaction and is
known as a straight shooter. Kashani-Poor told me that Weinberg
pulled my paper from his inside sport jacket pocket and said to the
group, “Have any of you seen this paper? It looks really interesting.”
This was especially vindicating since Weinberg wrote a seminal
masterpiece on the various problems with the cosmological constant
and the attempts to solve it. Weinberg’s work also provided concrete
criticisms and no-go theorems that ruled out many attempts to solve
the cosmological constant problem. Luckily my model was able to
evade Weinberg’s no-go theorem. My model is still a work in
progress, and my research group and I are still improving it to
confront how the cosmological constant is related to the onset of
dark energy today.4
This was my first step toward really working on quantum gravity.
Aside from the sort of aesthetic question I raised earlier in the book
about the big bang—why should gravity only be partially quantized?
—there are also pragmatic scientific reasons to seek a quantum
theory of gravity. Currently, there are a handful of unsolved issues at
the interface of quantum field theory and classical gravity. And many
of these problems find a home in the beginning and early evolution
of the universe when both gravitational and quantum physics are
expected to be active. These cosmological problems, such as the
origin of matter over antimatter, inflationary versus cyclic universes,
dark matter and dark energy, are so daunting that they compelled
me to deviate again and carry out research in both superstring
theory and loop quantum gravity. My collaborators and I have spent
the last two decades using these cosmic conundrums as a compass
to test and further develop theories of quantum gravity.
Both LQG and string theory have their own communities, and
there are strong feelings about the veracity of each approach. Both
have complementary strengths and weaknesses, ranging from
technical to conceptual challenges. My take is that both theories
provide tools and new concepts to address the unresolved problems
facing cosmology and particle physics, and both may not be
sufficient. It has always been my take to let the problems and
unexplained observations dictate which theories and tools to resort
to. Thus, I have had the fortune to use both LQG and string theory in
my research. While we have a handful of candidates for quantum
gravity, none are complete. It is my view that they are all parts of the
elephant, and we can use these approaches to teach us something
about what the final theory might look like.
11

A COSMOLOGIST’S VIEW OF A
QUANTUM ELEPHANT

We don’t have to get to some altered state of mind or enter a


wormhole to see how mysterious the universe is—it’s right in front of
our faces. Our electronic devices are controlled by quantum effects
of electrons, and yet, we still don’t have a clue about what the
ghostly electron is doing as it approaches the two open holes in the
double slit experiment. It does something, but whatever it does, it
seems like our mind’s eye is unable to visualize or conceptualize it.
The uncomfortable picture is that the electron behaves like a wave
when we are not looking at it, and not just a material wave but a
wave of potentiality—it’s as if some invisible puppet master is playing
tricks on us. If we try to interrogate the electron at the two open slits,
where we expect it to be doing its witchcraft, it goes back to
behaving like a particle, and the interference pattern is lost. Another
equally weird situation is when two spinning electrons are entangled
and separated at far distances. We still don’t know how one electron
can instantaneously know that the other’s spin was measured to
always have the opposite spin of its distant partner. Quantum theory
predicts this spooky action at a distance but refuses to tell us how it
happens. In other words, the math says that the spins should know
about each other at infinite distances, but the math doesn’t simply
provide a mechanism. These unresolved issues in quantum
foundations have led some physicists, including Einstein himself, to
believe that quantum mechanics is incomplete. However, as far as
precision experiments can tell, all the predictions of quantum
mechanics, no matter how bizarre, have been confirmed.
To me what is bizarre is how this real world, with planets, people,
and falling apples, emerged from an underlying virtual quantum
universe in the distant past. As I will soon discuss, we have strong
evidence that our entire universe in its infancy was quantum, and it
evolved to generate our macroscopic world. The paradox can be
simply stated as: What measured the wave function of the universe?
A first step to understanding the magical emergence of our classical
world is to ask how quantum mechanics and Einstein’s description of
gravity and space-time can be merged.
We discovered that when the gauge invariance principle is united
with special relativity and quantum mechanics, we have a pathway to
unify three of the four forces. We arrive at a unique quantum field
theory called the standard model that puts all three forces on the
same conceptual and mathematical footing. For decades physicists
have been challenged in uniting gravity with quantum mechanics.
Here is the elephant in the room: it is space-time that would be the
state that will exhibit superposition. So, what does it mean for a
region of space-time to have many superpositions? Despite this and
a handful of potential conceptual questions, there is good reason to
invite the quantum into space-time, since quantum mechanics has a
history of resolving the singularities and instabilities found in
classical physics. Classical general relativity, which is home to even
worse black hole and cosmic singularities, seems to be begging for a
quantum rescue. What is the culprit that makes quantum gravity so
difficult? We must go into this abyss as we expect the final theory of
quantum gravity, if it exists, to see what’s behind the black hole
singularity and what happened before and at the birth of our
universe. For me, quantum gravity holds the promise to give us a
new view of our universe far beyond our current imaginations.
Maybe we will be able to make sense of the hidden variables that
Einstein was after.
First, it’s useful to understand the two main doorways into
quantum gravity: background independence and emergence.
Background independence is a strong version of the principle of
invariance. Recall that the principle of invariance is a key concept in
all four forces, but it has a special role in general relativity: invariance
operates differently in gravity. The differences give us the first key as
to why quantum gravity differs from the successful quantization of
gauge field theories like the standard model.
We use coordinate systems to describe the position and motion of
an object, but according to the equivalence principle, the reality of
the object should not depend on the coordinate label given. We
could use any other coordinate system to describe the same particle.
Einstein realized that the laws of physics should not depend on the
choice of a coordinate system—this is the principle of general
covariance, an invariance principle for coordinate systems. The
invariance of coordinates is analogous to the name given to a
person. Suppose there is a person whose name is John. In Spanish,
his name is Juan. In Portuguese it is João. Regardless of the various
translations of the name, they are just labels for the actual person,
who is independent of the translations. So, if we can describe the
location of things with a coordinate system, then different
“translations” or choices of these coordinates are all equivalent ways
of describing the location of the thing. In principle, there is an
invariance encoding the infinite number of ways we could label the
position of an object at a point in space-time. We call this
diffeomorphism invariance, and it is at the heart of the relativity
principle.
Loop quantum gravity is a successful background-independent
formulation of quantum gravity. It was pioneered by Abhay Ashtekar,
Carlo Rovelli, and Lee Smolin, whose trick for quantizing gravity was
to reformulate general relativity so that the background
independence is formulated as a gauge theory (like the standard
model of particle physics) and to implement similar gauge theory
methods of quantizing as was done in the strong interactions of
quarks—that is, in quantum chromodynamics. A gauge field can be
quantized by having it wrap around a closed loop, like a lasso. The
wrapping of a concentrated amount of gauge field flux around a loop
is known as a Wilson loop, named after its inventor, Nobel laureate
Kenneth Wilson, and it was used to understand how quarks confine
in the strong interactions. We can think of Wilson loops as tubes that
carry electric flux—lines of a concentrated electric field. The idea is
that two quarks are attached to each other by a string of flux.
Similarly, gravity is equipped with its own gauge field called the
Ashtekar connection. This connection is the key ingredient behind
quantizing gravity, and it leads to the remarkable conclusion that the
basic building blocks of LQG are atoms of space-time called spin
networks, invented by Roger Penrose. These networks join to make
quanta of area and volume that have the size of the Planck length—
a minuscule distance defined by quantum theory as roughly .000000
000000000000000000000000000016 meters. If this is the case, then
it points to a way to tame the ultraviolet infinities found in quantum
field theory since quantized space has a minimal distance and
energy.
Another interesting feature of spin networks is the quantum
nature of time. When a spin network undergoes a quantum transition
it corresponds to a change in time, which ticks as a quantum by a
billionth of a trillionth of a trillionth of a trillionth of a second!
Since LQG is a fundamentally discrete structure, one of the big
questions is how to recover the smooth space-time that we see in
our macroscopic world—this is called the semiclassical limit. In
ordinary quantum mechanics the Planck constant is the parameter
that turns quantum effects on or off. So, by setting the Planck
constant to zero we expect to obtain classical physics. However, this
logic does not work in LQG, so some other scheme is needed to get
back classical space-time. There have been a handful of proposals
to get back classical general relativity from LQG, but it is still at the
level of current research. I think that if LQG is to be a successful
theory of quantum gravity it should not only tame the black hole and
cosmological singularities and recover the good predictions that we
already see in general relativity but also connect with some unsolved
problems in particle physics.
FIGURE 21: A spin-foam transition from one spin network to another
depicting quantized time steps.

One question my colleagues and I have been pursuing is how to


connect the classical limit of quantum gravity to be coincident with
particle physics. One exciting prospect is the realization that the
Ashtekar variable, a key ingredient in allowing classical gravity to
become LQG, enjoys a broken symmetry that is identical to the weak
force. Is this a coincidence? As I was reminded by my mentor, Nobel
laureate Leon Cooper, “If we knew what we were talking about, we
wouldn’t call it research.” I vividly remember the moment I realized
this coincidence between the weak force and gravity as a young
faculty member at the Institute for Gravitation and the Cosmos at
Penn State. It was actually a simpleminded insight that related a
coincidence that the symmetry group of the weak force and general
relativity, in the Ashtekar formalism, were the same, though these
two forces should have nothing to do with each other. But the
genesis of the intrigue occurred during my days at the Stanford
Linear Accelerator Center; I had always been puzzled about the
origin of parity violation in the weak interaction.
Parity, as we saw when discussing baryogenesis, is a symmetry
transformation that is analogous to looking in a mirror. When you
look at your left hand in the mirror it is identical but gets converted to
a right hand. Parity is a transformation that also inverts an object,
turning it upside down. For example, parity would transform a
spinning ice skater into an upside-down ice skater spinning in the
opposite direction. All the forces were thought to be invariant under
parity transformations. In other words, if we look at the scattering of
a left-handed quark with a left-spinning gluon field it would occur at
the same probability as a right-handed quark interacting with a right-
spinning gluon field. The big surprise was that the weak interaction
allowed interactions with one-handedness.
While it seems that the violation of parity is inconsequential, life
would not exist without it. The parity violation in the weak interaction
is at the heart of all nuclear decay processes as well as the chain of
subnuclear reactions necessary for fusion. The sun relies on the
nuclear fusion of two hydrogen atoms to make helium to provide the
energy necessary for life on Earth. So, there would be no life as we
know it without parity violation. Sheldon Glashow, Abdus Salam, and
Steven Weinberg were awarded the Nobel Prize for constructing the
correct theory of the weak interaction and uniting it with the
electromagnetic force. However, in their model of the weak force the
parity violation was assumed, and there was no deep reason as to
why parity violation emerged from some deeper theory and how it
would do so. My supposition from the resemblance of the Ashtekar
connection of general relativity was that parity violation in the weak
interaction came from quantum gravity itself.
So, what does one do when an idea appears that seems crazy
and too speculative? Well at the time, my strategy was to go to the
source, Abhay Ashtekar, the inventor of the Ashtekar variable.1
Ashtekar has a reputation for being an extremely technical and
detail-oriented kind of theoretical physicist, which he combines with
laser-sharp intuition. So, if you have an idea and take it to Ashtekar,
expect to walk out with your idea in a coffin to be buried in
mathematical soil. I knew of young physicists who would be very
careful to iron out any technical details before talking with the
master. But for some reason, Ashtekar always gave me a hall pass.
He treated me as if I had something to say even though my ideas
were often raw and inchoate. So, I felt comfortable telling him the
idea. After a few minutes, he told me that he didn’t notice this
connection between parity violation in gravity and the weak force and
encouraged and challenged me to pursue the idea. After months of
calculations and thought experiments, I developed a theoretical
model to explain the weak interactions parity violation as a
gravitational phenomenon.
One of the assumptions behind LQG is that the quantization of
gravity does not immediately take matter into account, especially
fermions. Recall that all known matter fields are fermions, and they
are special because they obey the Pauli exclusion principle.
Fermions have a geometric aspect to them since they also carry
information about the twistiness of space-time, a quantity called
torsion. In present versions of LQG, torsion is assumed to be
negligible—what physicists call vanishing. But if LQG is to properly
speak about fermions then it may have to relax the assumption of
vanishing torsion. Fermions are special because the reason your
hand does not go through a table is due to the exclusion principle,
which states that no two fermions can occupy the same quantum
state. So, macroscopic objects exist because fermions, like protons,
neutrons, and electrons, completely evade occupying the same
place. I believe that if LQG is to have a hope of having a
semiclassical limit, it will come from the simultaneous emergence of
the exclusion principle from a quantum foam where fermions are not
exclusionary. In other words, space-time becomes fuzzy when
fermions lose their special exclusionary powers, but some new, as
yet to be discovered, physics makes the fermions transition from
being nonexclusionary—“fuzzy”—to occupying different points in
space-time, linking the fermions’ distribution to the emergence of
classical space-time.
FIGURE 22: A representation of torsion in loop quantum gravity.

As I said, LQG isn’t the only way that I have tried to understand the
elephant of quantum gravity. Recall that our access to the quantum
world starts from first considering a classical system and applying
quantization to it. As in the case of the standard model and LQG, we
can quantize by writing down the relevant equations of the classical
system and employing the rules of quantization given by Paul Dirac
and Richard Feynman. In the case of string theory, which was first
proposed to explain certain patterns in the strong interactions, a
remarkable surprise presented itself. Take the description of a string
of energy that sweeps out a two-dimensional sheet in space-time.
We write a mathematical description of the energy for this string, and
we can quantize it the same way we quantize a particle. Naturally,
the quantum vibrations of the string come with a richer structure than
its point-particle cousin. Like a point particle the quantum string will
come in a superposition of vibrational states. However, the string will
also have other types of oscillations that are unique to its extended
nature.

FIGURE 23: Three representations of the key entities in string theory. To


the left, we have a point particle sweeping out a world line. In the middle,
we have a world sheet of an open string. To the right, we have a world
sheet of a closed string.

There are two types of string oscillations, massless and massive.


The big surprise to the quantum gravity researchers was that the
massless string oscillation includes a spin-2 particle. The only known
massless particle with spin 2 is the quanta of the gravitational field,
otherwise known as the graviton. This was the first indication that
string theory is not what the early string theorist ordered—a theory of
the strong interactions. String theory contains quantum gravity.
Even more remarkable is another stringy quantum effect. When a
string moves, it can emit a string that propagates and rejoins the
original string. This is known as a loop effect, which has a generic
effect in quantum field theories to change the strength of
interactions. In quantum electrodynamics, a loop effect changes the
strength of the interaction between electrons, and this effect has
been measured in the lab. In string theory, the same thing happens,
but, unlike particle field theories, an invariance that is unique to
strings is violated because of the loop effect.
String theory enjoys a symmetry that leaves the physics of the
string invariant when we zoom in on the string with a magnifying
glass. This symmetry, called conformal invariance, is analogous to
the image of a fractal, a pattern that repeats itself regardless of how
many times one zooms in or out of the pattern. This symmetry is
sacred for strings and must remain intact when the string becomes
quantized, otherwise the sacred conservation of probability is
destroyed, rendering the underlying quantum theory to make
nonsensical predictions. But when the quantum loop effect is
considered in string theory, this scaling symmetry appears to be
violated. Astonishingly, the loop effect in string theory generates
Einstein’s theory of general relativity so as to restore the hallowed
conformal symmetry of the string. Amazing! We start out with a
theory of a string with no gravity, turn on quantum effects, and
general relativity pops out. Master string theorist Edward Witten once
said that if we knew nothing about general relativity and knew string
theory, general relativity would be a prediction.
FIGURE 24: Schematic representation of gravity emerging from a string.

For string theory to be quantum mechanically consistent, the


resulting theory of gravity must live in ten dimensions. Many
physicists have lamented that we live in four dimensions, therefore
string theory makes the wrong prediction about the dimension of
space-time. However, the extra dimensions in string theory address
another mystery found in general relativity and the standard model:
What is the origin of the matter fields in our universe? One of
Einstein’s final efforts for a unified field theory was to extend the
notion of geometry of space-time for matter fields. He adopted an
idea first developed by Theodor Kaluza and Oskar Klein, who
showed in the 1920s that a five-dimensional theory of gravity seen
from a four-dimensional perspective is equivalent to Einstein’s theory
of gravity and electromagnetism. The electromagnetic field is
realized as the fifth dimension curled up in a small circle. How does
that work? The gauge field, whose quanta is the photon, contains all
the information about electromagnetism and arises from an
invariance of the photon moving on a circle, which is identified with
the symmetry of the extradimensional circle. This is a wild idea—that
space can cast itself as electromagnetic fields like a shadow in lower
dimensions. Of course, the question of what sets the size and
stabilizes the radius of this fifth dimension remains open, and we will
return to this issue when we discuss the role of string theory in the
early universe. Nevertheless, using the same logic as Kaluza and
Klein, we see that the ten-dimensional gravity theory in string theory
tells a similar story. The extra six dimensions will carry the
information about both the matter and the three force carriers of the
standard model.
Because common sense and experimental evidence so far both
point to a four-dimensional world, string theory must somehow
provide a way to understand why we cannot move through the extra
six dimensions and why they are not visible. One criterion is that the
extra dimensions are microscopically small. This condition leads to a
rich set of predictions. Like the Kaluza-Klein theory, the geometry of
the extra dimensions manifests itself as the fields of the standard
model. String theory starts to look like both a theory of quantum
gravity and a unified field theory. But there are some challenges that
have faced string theory for decades. First, on top of the fields of the
standard models, string theory also predicts a very large number of
other fields called moduli that have not been observed. These moduli
are fields that govern the sizes and shapes of the extra dimensions
that we do not observe in our low-energy world. String theory comes
very close to giving us the field contents and symmetries of the
standard model. Recall that the standard model did not give us any
information about the coupling constants and masses of the
particles. Fortunately, string theory moduli can set the values of
these parameters. Since these moduli are dynamical fields in string
theory, then what kind of physical mechanism fixes them to the
values that are observed in the standard model? If we are to believe
that string theory is the answer to reality, then what became of these
uncountably large moduli fields?
What I personally find very promising about string theory, despite
its issues in giving us a realization of our four-dimensional world, is
that it naturally encodes the emergence principle. As we saw earlier,
we start with a theory of strings with no gravity and it spits out the
equations of general relativity and gauge theories (albeit in ten
dimensions) when the string experiences quantum corrections. Also,
there are realizations in string theory (and loop quantum gravity)
from which gravitational theories emerge holographically.
Holography, invented by Gerard ‘t Hooft and Leonard Susskind, is
the idea that physics is encoded in a space that is one dimension
less than what is experienced. According to the holographic
principle, our experience of this four-dimensional world is actually an
illusion, and the information is encoded in a three-dimensional
screen that lives at the boundary of the four-dimensional world. A
good geometric realization of this is to imagine that our world is a
three-dimensional, solid ball. Then by holography all the information
about the ball is actually encoded in the two-dimensional surface of
the ball, not in its volume. Exactly how the three-dimensional world is
reconstructed depends on the exact nature of the model at hand.
Hints of the holographic principle were inspired by laws of black hole
mechanics discovered by Jacob Bekenstein and Stephen Hawking.
They found that for a black hole the entropy did not depend on the
volume within the black hole but the area of the horizon. The
information of the relevant degrees of freedom was found to be
residing on the two-dimensional surface of the black hole and did not
care about its interior.
String theory realizes holography by using the correspondence
principle that Niels Bohr argued for in quantum mechanics. In string
theory there are two theories that are complementary, or dual, to
each other. One theory is a nongravitational conformal field theory
(CFT), actually a theory of gauge fields, like the ones found in the
standard model but with a lot more symmetries. This gauge theory is
complementary to a theory of gravity with maximal symmetry with a
negative cosmological constant, known as AdS. Since its inception
the duality has passed every nontrivial test of the equivalence
between these two theories. Taken literally, this means that gravity
emerges from the dynamics of the CFT—which itself has no gravity.
The string theoretic realization of holography was invented by the
Argentinian physicist Juan Maldacena. I remember when
Maldacena, at the time a postdoc at Princeton, wrote this first
groundbreaking paper in 1997 now referred to as the AdS/CFT
correspondence. The AdS/CFT duality relates a gravitational theory
to a space-time called AdS, which is a homogenous and isotropic
space-time with a negative cosmological constant. I was a graduate
student and wondered how he set up the string theory calculation to
realize the holographic duality. The paper quickly got international
acclaim, and up to today its impact and number of citations by others
have not lost steam.
Whatever quantum gravity will look like, I’m convinced that
holography will play a key role.
12

THE COSMIC BIOSPHERE

The total number of minds in the universe is one.


—ERWIN SCHRÖDINGER

I have always wanted to know the answer to three questions: Why is


the universe expanding? Is there life elsewhere in the universe? And
if so, how did it originate? These might seem like distinct concerns,
but I have come to the opinion that the questions about the
expanding universe and life are linked. The last several chapters
have focused on the first question. Now I want to deal with the
second.
The thing is, I don’t think they are first and second questions.
While we don’t have complete answers to these questions, I think
origination and persistence of living systems are traced back to the
expansion of the universe. You should be at least astonished by this
claim, so let’s dive into it.
When we look at cosmic history, we see the chain of events that
unfolded to generate the large-scale structure in the universe.
Cosmic time unfolded with primordial quantum vibrations that
blossomed into a hierarchy of stellar environments called galaxies
and planets that are now part of an interwoven cosmic structure.
When we look at the large-scale structure of galaxies, we assume
that life and the universe that it inhabits, borrowing from the famed
biologist Stephen Jay Gould, are “non-overlapping magisteria.” To
most cosmologists, complex systems like life are of little
consequence to the problems we are trying to solve, such as the big
bang singularity and the parameters of the standard model. To my
biologist friends, life is housed in a biosphere that is decoupled from
the happenings of the universe out there. But what evidence do we
have that life and the universe are truly decoupled? Conversely,
what evidence do we have that life and the universe are coupled?
My dance with biology and physics started way back in college,
when I considered majoring in biology. I have never been able to
shake the biology bug, and my current research in cosmology has
reawakened my biological questions, even though I put formal plans
aside for a while. At my college, bio majors were required to have a
year of organic chemistry, so laziness got the best of me. Still, my
musings in biology persisted throughout my studies in physics. In the
middle of my second year of graduate school I decided that I wanted
a change, so I approached my field theory professor Gerry Guralnik,
a codiscoverer of the Higgs boson, and confided to him that I was
interested in biophysics and was considering leaving physics
altogether. Then Guralnik said, “Let me call my former PhD adviser
Wally now and get his advice.”
The Wally that Gerry was referring to is Walter Gilbert, a
theoretical particle physicist who had also caught the biology bug
and ended up winning the Nobel Prize for a key discovery in
genetics—the nature of stretches of DNA known as introns and
exons—which, among other things, led to the human genome project
and gene therapy. Guralnik made the call and within minutes Gilbert
invited me to visit his lab at Harvard. I spent three hours talking with
Gilbert and he gave me an in-depth tour of his lab. He also made a
recommendation for me to work in Harvard’s biophysics program.
That summer, thanks to Gilbert’s recommendation, the renowned
biophysicist Jim Hogle offered me a job in atomic resolution virus
structure determination using X-ray crystallography. From that
experience, I learned some new tools and gained a deep
appreciation of the complexities involved in biology and the
inapplicability of physical reductionism in attempting to comprehend
life’s processes. It also comforted me to learn that I wasn’t alone:
other physicists and mathematicians—including many of the greatest
—had explored their own biological musings. People including John
von Neumann, Eugene Wigner, Claude Shannon, Norbert Wiener,
and Roger Penrose, to name a few. I especially think that we can
take some lessons and inspiration from the story of Erwin
Schrödinger, whose audacious speculations and predictions in
biology have been hugely influential. Physics has undergone
exponential growth since Schrödinger’s time, and in this “naive”
spirit, I will playfully venture into ways in which biology may inform
some mysteries in physics and vice versa. But first let’s learn from
Schrödinger’s scientific audacity and ingenious contributions to
biology.
Schrödinger moved to Ireland to escape the Nazis in 1938, and
continued his work in Dublin. In 1943 he gave a series of lectures at
Trinity College that would eventually revolutionize biology. In 1944
they were published in a tiny book called What Is Life? In it,
Schrödinger speculated on how physics can synchronize with
biology and chemistry to explain how life can emerge from inanimate
matter. Schrödinger proposed his central question by asking, “How
can the events in space and time which take place within the spatial
boundary of a living organism be accounted for by physics and
chemistry?”
In other words, how is it that the same laws of physics that
describe a star account for the intricate processes of metabolism
within the living cell? Schrödinger quickly admits that the physics of
his time was insufficient to explain some of the ingenious
experimental findings that his predecessors and contemporaries had
made about the living cell. He even considered the possibility that
physics as it was known then may not be enough, when he said,
“One must be prepared to find a new physical law prevailing in it.”
Despite this, Schrödinger marches ahead, using the physics of his
time to make some auspicious predictions that inspired the discovery
of the double helix structure and functioning of DNA.
Quantum mechanics plays a key role in life, since it is the
quantum that is necessary for the stability of atoms and the bonding
rules to construct the plethora of molecules found in living matter.
(As to whether the more “exotic” quantum properties, such as
entanglement or quantum tunneling—where a wave function can
actually pass through a barrier—may play a role in life, this remains
an open research issue. We’ll get to that.) Schrödinger opens his
argument by conjuring quantum mechanics as the starting point to
understand the difference between nonliving and living matter. For
example, the bulk properties of a piece of metal, such as its rigidity
and ability to conduct electrons, require an emergent long-range
order, as we saw when I first raised the emergence principle. These
properties should be a result of the bonding mechanism and the
collective effects of the quantum wavelike properties of electrons in
the metal’s atoms. Schrödinger then describes how the atoms in
inanimate matter can organize themselves spatially in a periodic
crystal, before making a daring leap. Life clearly is more complicated
and variable than a piece of metal, so periodicity isn’t going to cut it.
So Schrödinger makes a bold proposal: that some key processes in
living matter should be governed by aperiodic crystals. More
astonishing, Schrödinger postulates this nonrepetitive molecular
structure—which will turn out to be a great description of DNA—
should house a “code-script” that would give rise to “the entire
pattern of the individual’s future development and of its functioning in
the mature state.”
Before Schrödinger’s time biologists had the idea of the gene, but
it was a formless unit of inheritance, with much that was left
unknown. When Schrödinger proposed his idea about how genetic
material should work, it was completely unanticipated in anyone
else’s work. Today, of course, the idea that genes are governed by a
code, similar to a computer code, which could program the
structures and mechanisms of the cell and determine the fate of
living organisms, might just seem to be common sense. While I say
this, exactly how this is accomplished at a molecular level is still a
rich research enterprise in biology. What is remarkable is that
Schrödinger used reasoning stemming from quantum mechanics to
formulate his hypothesis. Schrödinger was an outsider to biology,
and this naturally made him a hidden deviant and ripe for making a
paradigm-shifting contribution. Whether quantum mechanics was
operating in some subtle way for life’s processes was not central to
Schrödinger’s argument, but his new line of reasoning provided a
completely new set of concepts to explore novel mechanisms. It’s
worth noting that no one took Schrödinger’s “code-script” seriously,
even after biologist Oswald Avery’s 1944 publication of a paper that
gave strong evidence for DNA as the carrier of genetic material. Of
course, Avery’s work itself wasn’t immediately embraced: as
described by science historian Matthew Cobb, part of the reason for
this lack of excitement over Avery’s discovery was that DNA was
thought to be a “boring” molecule with a repetitive structure—exactly
what Schrödinger had said a gene could not be. Nevertheless,
Schrödinger’s quantum reasoning led to a prediction that the
aperiodic structure of DNA carried a code that could program life.
Schrödinger’s unappreciated insight came about from an
auspicious set of events. In 1943 the US Scientific Research and
Development Committee hired some scientists to study information
from radar for antiaircraft purposes. Among the scientists were
Claude Shannon, the pioneer of information theory, and Norbert
Wiener, who found connections between control systems in
machines and biological life and coined his findings cybernetics.
Both fields have surged into prominence today, especially in the form
of quantum information theory, as well as machine learning and
artificial intelligence, but they were hugely influential in our
understanding of the theoretical and computational underpinnings of
life.
As important as Shannon and Wiener are, the game really
changed when computer science pioneer John von Neumann
argued that a gene was the carrier of information. Von Neumann
imagined a gene as a tape that could program an organism. He
made an analogy between self-replicating machines and cellular
replications. In this case, in order for a machine to replicate itself,
there must be an underlying mechanism present to copy the
information that specifies the machine itself. In a 1951 conference
proceeding, the godfather of developmental biology Sydney Brenner
said “[Von Neumann] divided the machine—the automaton as he
called it—into three components: the functional part of the
automaton; a decoding section which actually takes a tape, reads
the instructions and builds the automaton; and a device that takes a
copy of this tape and inserts it into the new automaton.” It wasn’t
immediately obvious what von Neumann was onto, at least not to
Brenner: “I think that because of the cultural differences between
most biologists on the one hand, and physicists and mathematicians
on the other, it had absolutely no impact at all. Of course, I wasn’t
smart enough to really see then that this is what DNA and the
genetic code was all about.”1 Brenner’s observations about how
cultural differences between biologists, physicists, and
mathematicians were in part responsible for biologists missing von
Neumann’s earth-shattering insight about information, genes, and
replication remain relevant today. The silos of scientific disciplines
and scientific social orders still limit how scientists work.
There were two more insights that Schrödinger posited between
life and physics. I will leave it for you to read his landmark book What
Is Life? about his reasoning that living things have to negate the
second law of thermodynamics, which describes the fact that closed
systems evolve to maximize disorder, or entropy. Schrödinger called
this negentropy: “What an organism feeds upon is negative entropy.
Or, to put it less paradoxically, the essential thing in metabolism is
that the organism succeeds in freeing itself from all the entropy it
cannot help producing while alive.” Living entities have to stabilize
their structure and function over their lifetime against the
fundamental tendency for disorder, since entropy always increases.
Other biologists, mathematicians, and physicists have further
developed Schrödinger’s idea into rigorous mathematical statements
about how this works.
His third insight has to do with consciousness, and we’ll come
back to that in the conclusion of the book. First I want to ask, how
would Schrödinger revise What Is Life? with the new developments
in modern biology and cosmology?

Over the years I have developed a pattern: I have chosen to make


friends with biologists. Having those friendships means that I’m
always asking about what they think is cool in biology. In the past
several years I’ve had frequent conversations with my friend
Salvador Almagro-Moreno, a molecular biologist. Sometimes
meeting over a drink, we exchange ideas: I tell him my latest
musings in cosmology research and he tells me about his musings in
biology. It probably won’t surprise you to learn that Salvador is a
proud owner of a first edition of Schrödinger’s What Is Life? He and I
share a vision that I think many of our fellow biologists and physicists
would find too deviant, and even repugnant. We don’t let that stop
us, and there have been many times that we stayed up late talking
about this fascination. There was definitely some strategy here: in
part our conversations were an exercise in deliberately generating
an outsider perspective, hopefully to benefit each other’s research
insights. But it wasn’t all so calculating: it was a lot of fun, and just
one of the things that make our jobs as scientists so rewarding. What
is interesting to me about our discussions is not just the ideas
themselves, but how our conversations generated new questions in
both our zones of inquiry.
In 2014, I held the E. E. Just chair in natural sciences at
Dartmouth College. Ernest Everett Just was a pioneering Black
developmental biologist who cemented the early field of epigenetics.
His remarkable life story, scientific contributions, and legacy are
majestically described in Black Apollo of Science by Kenneth
Manning. As the E. E. Just professor, I had the responsibility of
running a university-wide science program for promising students
from underrepresented groups interested in science. Salvador was
awarded the E. E. Just postdoctoral fellowship and helped me with
the program by leading its mentoring and teaching component.
Nowadays, Salvador is a biology professor at the University of
Central Florida, one of the largest Hispanic-serving institutes in the
United States, and is leading similar efforts as the E. E. Just
program. One day we were discussing some program business
when the conversation, as usual for us, evolved. I started telling my
friend about a research idea I had in cosmology. The idea was a new
mechanism to understand what is called the fine-tuning problem: as
with the value of the cosmological constant, there are other
constants of nature that have values that are just the way they need
to be so that life could happen. A hugely important set of constants,
known as coupling constants, that determine the strength of the
force—such as electromagnetism or the weak force—are involved in
interaction between particles. The idea I told Salvador about was
that, if the universe went through a large number of cycles of
collapse and expansion, the big bang phase would provide an
opportunity to reset the values of the coupling constants. Most cycles
of expansion might not host life, but eventually the universe would hit
the jackpot. Salvador asked me very precise questions about my
project, and the next day, in a state of elation, he told me that he was
able to implement the cyclic-universe idea into the development of a
theory based on some experiments he completed involving genetic
evolution in biology.
One of Salvador’s concerns was understanding what it took for
bacteria to become pathogenic—capable of causing disease in a
host like us. According to him, various combinations of genes, like a
slot machine, can hit the jackpot in giving the bacteria the ability to
become harmful. Salvador made a brilliant analogy between the
coupling constants in my cosmological model and the various forms
of a gene, known as alleles of the gene. He likened the bacterium’s
population cycles and environmental factors to a cyclic universe as a
mechanism to change the different combination of genes. The
different replicating populations represented the many cycles in the
model, each with its own alleles or variations. This led him to
propose the theory of virulence adaptive polymorphisms (VAPs).2
The work was published in the top journal in microbiology. The idea
worked and has made a big mark on his field.
Now it was time for my friend to return the favor and inspire some
ideas for cosmology. Over the years we had mischievously
developed a conviction and intuition that there is a hidden
interdependence between living systems and cosmology. I came at
the question from cosmology. Salvador came at it from a biological
direction. And to our delight this question led us to a few issues that
the big bang and living systems (such as ecosystems) have in
common.
A major concern for both of us has to do with the flow of entropy
in the universe, whether at biological or cosmological scales. In the
epoch in the early universe before there were stars and planets, the
universe was mostly filled with an equal amount of radiation and
matter, where the photons and electrons were in thermal equilibrium.
If a collection of gas molecules occupy a closed system, say a room,
they are going to tend to thermal equilibrium—the temperature,
essentially, will become the same throughout the room. As they
approach equilibrium, their entropy will increase. Entropy is a
measure of disorder, and so of ignorance about what the gas
molecules are doing. The entropy will increase when there are more
particles and more space for the particles to potentially occupy—it
becomes increasingly difficult to specify what they are all doing.
Entropy also negatively affects the ability of any system to do work.
Physics, biology, and chemistry rely on an important concept called
free energy, which is a measure of how much energy in a system
(such as a living cell) is able to be used for work. Mathematically we
can express free energy as ΔF=ΔE−TΔS. The equation states that a
system can do work with a positive change in free energy (ΔF),
where a positive contribution comes from a change in energy (ΔE)
and a negative contribution from the change in entropy (ΔS). The
entropy contribution to free energy reduces the amount of energy
that can be used to do work. For example, sunlight shining on Earth
generates free energy, which we calculate by adding the contribution
of the potential energy stored in the wavelength of the photons and
subtracting the entropy from the array of photons.
But there are some important caveats having to do with gravity.
The gravitational expansion of the universe keeps the matter and
photons in the cosmic microwave background homogeneously
distributed. Situations of extremely high gravitational entropy are
contained when matter collapses into localized objects like black
holes, which did not exist in the era when the CMB formed. So, when
the universe expands, gravity acts to distribute matter in a
homogenous, ordered fashion, and this lowers the entropy of the
universe. When gravity acts to coalesce matter into supermassive
black holes, entropy goes up: the heavier the black holes the larger
the entropy.3
And there is an important problem. The photons and matter in our
universe were in equilibrium during the CMB epoch, so the entropy
then was high. But as the universe continued to evolve, the entropy
would continue to increase. This implies that the entropy of the very
early universe, before the CMB epoch, must have been very low.
At the largest observable distances, we see a connected pattern
of large-scale structure of galaxies distributed across the universe.
As the universe continued to expand and cool, out-of-equilibrium
structures with varying complexity like stars, clusters of galaxies, and
life formed. The structures will contain lower entropy than the rest of
the universe. By starting off with low entropy, the universe is able to
arrest the growth of entropy against the trend of the second law, by
concentrating regions of lower entropy within cosmic structures.
These cosmic structures, such as stars, store potential energy; in the
case of stars, it’s from the rest mass of hydrogen, which will release
highly energetic photons from nuclear fusion. This entropy-lowering
network of structures becomes the main currency for the biosphere
and life on planets. Even the father of thermodynamics, Ludwig
Boltzmann, said, “The general struggle for existence of animate
beings is therefore not a struggle for raw materials… nor for energy
which exists in plenty in any body in the form of heat, but a struggle
for entropy, which becomes available through the transition of energy
from the hot sun to the cold earth.”4 Nevertheless, even as the
universe deviated from homogeneity, by seeding and forming lower
entropy structure, entropy elsewhere in the universe continued to
grow. And entropy also has a tendency to grow within those
structures. This makes entropy, or its absence, a key player in
sustaining cosmic structure, such as stars and life; therefore, an
early lifeless universe with low entropy is necessary for life here on
Earth. Stars like our sun radiate free energy to the earth. This free
energy is absorbed by electrons in plants and used for the necessary
chemical work for its living function. The plant will release this energy
in the form of heat and give off to the universe more entropy than it
took in.
Unfortunately, it is difficult to explain with our current
understanding of physics why the entropy was so low in the early
universe. In fact, this problem of the low entropy we demand of the
big bang is one of the major problems with the theory. It was first
identified by Roger Penrose. Its solution remains a mystery.5 I
remember discussing this problem with Penrose on a nice summer
walk, and we both agreed that if we wanted to understand how
gravity could have helped set up this unlikely scenario, we were
going to need the real connection between entropy and gravity,
which is currently lacking, to reveal its nature. One hint is that at the
earliest moments of the universe, close to the big bang, the
curvature of space-time approached infinity. Whatever new physics
tamed this infinity should tell us why the entropy of the universe was
so low. We will get to this.
The biology side of the story stems from Salvador Almagro-
Moreno’s research into the genetic and ecological drivers that lead
populations of harmless bacteria to evolve and emerge as
pathogens. Crucial to the story is that it isn’t just a question of the
genetic code of the bacteria. One of Salvador’s mantras is that life is
an adaptive phenomena responding to constant and unexpected
changes in pressures from the environment. If life can have more
channels and resources for being adaptative, it will find a way to use
them. Central to his research is understanding evolution from the
genetic code in a population of organisms, and the epigenetic
influences from the ecosystem. Epigenetic factors are called that
because they sit on top of genetic factors, and they are one of those
other channels for adapting beyond changes to the genetic code. For
example, an environmental factor, such as a pattern of electrical
current hitting the cell membrane during replication, can enhance or
suppress certain genetic factors, leading to completely different
features in the phenotype of the offspring.
FIGURE 25: A simulation of the large-scale structure of the universe.
Each dot represents a galactic system.

This makes an organism an emergent phenomenon, where the


final shape of it is not contained in the individual pieces and
influences that make it up. Recall that in emergent phenomenon in
general, it is the interactions of the building blocks that collectively
exhibit the emergence. This also implies that a population is
emergent, too. Living things comprise a network of interactions that
is mediated through the environment. A living system is able to
regulate billions of cells to maintain its overall functioning. Beyond
that, collections of organisms belong to a network called an
ecosystem, which also maintains a dynamical equilibrium.
This extends all the way to networks at life’s largest scales. The
idea of the earth being a self-regulating ecosystem was
codiscovered by James Lovelock and Lynn Margulis in the 1970s,
and it became known as the Gaia theory. The name Gaia came from
the goddess who personified Earth in Greek mythology. In response
to the name, Lovelock, a chemist, observed that “biologists scorned
it… it gradually became known as Earth Systems Science, but it is
the same thing.” Whatever you call it, the takeaway for me is that the
flow of negative entropy exists not only for individual living things but
for the entire earth. The sun sends free energy to the earth, and
through a chain of complex interactions, the energy gets distributed
through a network of interactions to living things, each relying on it to
maintain its complexity in the face of increasing disorder. But there’s
no free lunch: when living things release this energy back into the
environment, they mostly do so in a form that has higher entropy
than what they received. Salvador and I noticed the uncanny
parallels between living systems and the evolution of the universe
through the lens of entropy.
This could seem like a coincidence in the behavior of the universe
and of life, but we decided to treat the parallel as though it were not.
Instead, we proposed that Schrödinger’s idea of negentropy is one
of the central organizing principles of the evolution of the cosmos
and the existence of life. Salvador elected to call this the
entropocentric principle, a wink at the anthropic principle that first
emerged from string theory and caused such a controversy when I
was first working on the vacuum-energy problem. The anthropic
principle, in its strong form, states that the universe is fine-tuned for
life. The laws of nature and values of coupling constants of their
interactions have the values that are consistent with life on Earth.
For example, if the strength of the nuclear interactions differed by a
few percent then stars would not be able to produce carbon and
there would be no carbon-based life. The fine-tuning problem may
not be as severe as it seems. In research I conducted with my
colleagues Fred Adams, Evan Grohs, and Laura Mersini-Houghton,
we showed that the universe can be fit for life even when we let the
constants of nature like gravity, vacuum, and electromagnetism vary,
so long as they vary simultaneously.6 Maybe we don’t need the
anthropic principle after all. The entropocentric principle, on the other
hand, is harder to shake. If the universe was unable to provide
pathways that enabled it to transfer regions of lower entropy, then life
as we know it would not exist. We call this biological dependence on
the entropic relationship of the cosmic structure the entropocentric
universe. Living systems situate themselves to reduce their entropy
by expelling it out into the environment, while consuming energy
from their environment. Did the universe play the same game near
the big bang?
13

DARK IDEAS ON ALIEN LIFE

Salvador Almagro-Moreno isn’t the only person with whom I went


exploring in new territory of the mind. When I lived in the Bay Area, I
used to get together with my friend Jaron Lanier to explore the
implications of spectacularly weird thought experiments.
Occasionally, one of these conversations would lead to an
interesting outcome. This chapter explores one of them. Outlandish
thought experiments have been essential in the intellectual history of
science, but the point isn’t the weirdness itself. The payoff of thinking
about strange things like Schrödinger’s cat, the infamous cat that is
alive and dead at the same time, is not necessarily that we should
then “believe” in the existence of such a cat. Instead, we can hope
that uncommon ideas will shed light on the murky margins of our
thoughts; in the case of Schrödinger’s cat, in dealing with the
question of superposition. The point is not to confuse or bamboozle
people, but to eventually find a way to think that makes more sense
and is a little less murky.
The bizarre notion I want to consider here came from a
discussion of the search for alien life forms. There are a variety of
ways to look for signs of alien life in the universe, usually involving a
large array of telescopes. One approach is founded on the hope that
perhaps astronomers will get lucky and chance upon an alien radio
broadcast. But in the thought experiment Lanier and I explored, we
considered a different and far more dramatic possibility. Suppose
that there are lots of alien civilizations running hugely capacious
quantum computers of the sort that Google and others are just
beginning to build here on Earth. This leads to a question of high
weirdness: Would an extreme amount of very distant quantum
computation result in any astronomically observable effect? Could
we humans see evidence of a universe teeming with quantum
computers by carefully examining the night sky?
We thought about various ways this might be possible, but in the
end we focused on one wonderful possibility. So here it is: First, alien
quantum computers could explain the mystery of dark energy,
because computation by multitudes of alien creatures across the
universe bends (or rather unbends) the universe as a whole.
Because we can observe the effect of dark energy, accelerating the
expansion of the universe, this implies that we have already seen
evidence that our universe is alive beyond us—we just haven’t
recognized it as such! And we found, fortunately, that contemplating
this almost imponderable notion has a human-scale practical payoff:
it helps us clarify how we think about plausible relationships between
gravity and quantum information. (If you think this is strange, you
should read some of the competing ideas. One recent paper
suggests that dark energy is actually a sign that time is about to
cease to be time and turn into space instead. We’d then be frozen
out of time, but be four dimensional. Compared to that, our proposal,
aliens and all, is practically tame.)
Let’s go through the argument step by step: What is a quantum
computer and why would aliens be using them? Let’s assume that,
just like us, plenty of alien civilizations will want the best possible
computers for some purpose or other. For the sake of argument,
we’ll assume the aliens want to enjoy high-quality virtual reality, and
so they build computers to make that happen.
If the computers that run alien VR are of the classical kind we use
these days (based on the mathematical framework laid down in the
mid-twentieth century by computer science pioneers John von
Neumann and Alan Turing), then aliens would generally endure an
inferior sort of virtual-world experience. You might think that classical
computers should be up to the task—after all, the special effects in
movies are getting fairly realistic, and classical computers are able to
calculate those effects—but they are not. Remember that movies are
prepared in advance. Virtual reality, however, must create sensations
for the human body on the fly and as quickly as reality does.
Classical computers can’t work that fast. Furthermore, there are
cases where the human body is able to respond to reality at the
highest possible level of sensitivity. For example, the retina can, in
certain cases, generate a neural response to a single photon. In a
case like that, the human body has become as discriminating as
physics can possibly allow. Just as classical computers can’t be as
fast as the universe, they can’t be that discerning, either. If we
assume that aliens elsewhere also evolved to be as sensitive to this
ultimate, quantum level of reality in some special cases as we are to
light, then when they try to design a nonquantum supercomputer and
VR apparatus that could simulate reality at the ultimate level of
detail, they would have run into problems. That’s one reason we
guess that discerning aliens would seek the power of quantum
computers to run their virtual worlds.
Quantum computers are not yet adequately developed for
practical uses on Earth, but they have the theoretical potential to
pulverize regular computers in a wide range of calculation contests.
A quantum computer can work as if there were copies of it in many
parallel universes at once, simultaneously exploring variations of the
computational task at hand.
Suppose the computer is calculating what a virtual rose petal
should feel like to your fingertip. The rose petal is pliant, so every
part you touch changes all the other parts you touch. You have to
calculate all the parts at once, and there’s only a single solution that
consistently reconciles all the tiny events in different locations of the
petal so that it feels realistic. A quantum computer can be calculating
a huge variety of different versions of the petal simultaneously, even
though only one variation is the correct version. That correct version
can then be instantly presented to your fingertip, perhaps by the
“octopus butler robot” that Lanier has imagined in his book Dawn of
the New Everything, as if the computer had somehow known which
variation would be correct from the start.
One big engineering problem for quantum computer designers is
heat. Heat is an almost universal problem for any computer designer.
Every time you change a bit inside a computer, you’re doing at least
a little bit of work, whether that bit is implemented as a bead in an
abacus or as a charge in a semiconductor in a silicon chip. Work
always gives off heat.
Let’s consider the example of the abacus. When you move a
bead up and down you generate some heat from friction on the wire
the bead is sliding on. If you do this only a few times a second, you
won’t even notice that heat, but if you move the bead millions of
times a second you will melt the wire. Now consider a quantum
abacus. This would be a little like having a bunch of copies of the
abacus in different parallel realities, each with the beads in different
positions, each exploring a different variation of a problem. You can
think of each individual bead as being like Schrödinger’s cat: you
can either think of a bead as being both up and down at the same
time, or that in each particular universe it is either up or down. We
call this kind of bead a qbit (quantum bit) instead of a bit. If the
quantum abacus gets hot, the beads start jiggling, in just the way
that oil in a hot frying pan will start sizzling. If the beads jiggle too
much, it becomes impossible to say which universe has a bead that
is up or down, which means that the differences between the states
of the beads in the parallel universes disappear. When that happens,
the quantum advantage also disappears.
The role of the beads in a quantum abacus is usually performed
by fundamental particles, which means that even a tiny amount of
heat will have this ruinous effect. Quantum computer designers here
on Earth have been forced to advance the state of the art of extreme
refrigeration. The machines have to be run at ultracold temperatures.
There are other ways to deal with the issue, however. One
potential way to reduce the heat problem is with a topological design.
Topology is a field of mathematics that describes how things
connect. For instance, if you put a bead on a circular wire, the rules
of the game change because pushing up on a bead that is already
up will force it around to become down. The meaning of an “up push”
becomes dependent on the previous push. When you change the
way things connect, you change the kinds of information that can be
contained in those things. An example from human cultural history is
that the ancient Incas used elaborate knots called khipu as a record-
keeping scheme and a form of written language.
If you use topology to hold information, you don’t have to worry
about heat quite as much as you do when you are moving bits
around within a fixed topology. Instead of just moving a bead, for
example, you could also bend and change the connections of the
wire so that the bead moves in different ways. In addition to loops,
you might explore various knots, branching structures, and so on.
Enough heat could still melt the wire and ruin your topology, but if the
heat is only enough to jiggle the beads, then the topology won’t be
harmed and you can preserve a lot of information.
Physicists like Nobel laureate Frank Wilczek are trying to figure
out how to make a practical topological quantum computer on Earth.
The general idea is to move around artificial, flat, fundamental
particles called anyons (which we first encountered in Jim Gates’s
work) within tiny confines so as to tie knots with the paths that they
trace as they move. This is a remarkable idea because the knots are
only knots if you think of time as one of the dimensions within which
the strands are held tight.
So far, so good: none of these ideas about quantum computers is
anything other than mainstream, not radical. But Lanier and I went a
step further to propose that alien computers are not only topological
quantum machines, but that they are gravitational. The particles they
tie knots with are gravitons. And this takes quantum computers into
the realm of dark energy.
As we have discussed, dark energy is treated as synonymous
with the idea of the cosmological constant these days. That wasn’t
always the case. When Einstein formulated the theory of general
relativity in 1915, the math predicted that the universe should be
expanding. That seemed wrong, so Einstein added the cosmological
constant so the universe would be static. Then in 1925 Edwin
Hubble gathered data that showed that the universe actually was
expanding, so Einstein got rid of the constant. Then, in 1998, our
friend Berkeley astronomer Saul Perlmutter gathered even better
data and showed that not only is the universe expanding, but the
rate of expansion is accelerating. And so physicists brought the
cosmological constant back, but with a new value. And that’s what
we call dark energy, and in case you forgot, it’s weird: it behaves like
a repulsive fluid with negative pressure that fills all the space around
us, pulling everything apart. Given a choice, it doesn’t sound like the
sort of fluid you’d want to swim in, but we are swimming in it.
As we have seen, at present there is no solution to the dark
energy problem. Prominent physicists like Ed Witten have called it
the greatest embarrassment of theoretical physics. It has forced
physicists back to the drawing board—all the way back to the
foundations of quantum mechanics and general relativity. And for me
and Jaron Lanier, that meant quantum computers.
Before we get to the alien computers, however, I want to consider
one last difference between Einstein’s general relativity and quantum
field theory that has kept physicists from reconciling them. That is
the passage of time. In general relativity there is no master clock, but
in quantum field theory time is expressed using a master clock that
sort of lives off to the side in a metaphysical realm. There are certain
phenomena, like black holes and the vacuum, which can be
approached by either theory. In the case of the vacuum, we can
clearly see some ways in which quantum field theory is unsatisfying.
For instance, remember that general relativity is all about curved
space-time. In fact, the cosmological constant is just an adjustment
of the curvature. It turns out that a detailed mathematical analysis
reveals that quantum fields in the curved spaces of general relativity
don’t have unique quantum states. An example of this strange state
of affairs is called the Unruh effect, named after Bill Unruh, a
physicist who described it, which predicts that the vacuum will
appear hotter to an accelerating observer. If the effect is eventually
observed, then most of what we write here will be rendered wrong,
but for the purposes of our argument we’re assuming the prediction
is wrong and that the nonrelativistic inner self of quantum field theory
is at the root of the mistake.
Recall one of the usual images invoked to describe general
relativity: if you try to use a passing train naively as a pitch pipe,
you’ll inevitably sing out of tune. This is because the train’s whistle
seems to go up and down in pitch as it flies by. What’s really going
on is that sound waves wash over you in different concentrations
depending on where they were emitted as the train travels relative to
you. When they get bunched up, because the train is approaching,
they seem to be higher in pitch. That creates an illusion of a
changing tone when the train goes by.
In the same way, quantum field theory asks a traveler in an
accelerating spaceship to consider an absolute clock, but that clock’s
time cannot be taken at face value, just like the sound of the moving
train whistle can’t be used as a steady reference. Screwing up the
measurement of time screws up everything else. The vacuum should
seem hot, according to quantum field theory, for the same reason the
whistle seems to get higher. Relativity doesn’t make this mistake, but
quantum mechanics, the grandfather of quantum field theory, has
passed on this “bad gene.” Within the context of quantum field
theory, we don’t have a truly relativistic platform with which to
describe the vacuum.
Back to the possibility of gravitonic topological alien computers!
Recall the earlier image of the many copies of a quantum abacus,
distributed across many realities, all working in parallel. When we
move to topological abacuses, where the wires can be tied into
loops, trees, knots, and so on, what is going on in all those parallel
universes? There is an interesting divergence. On one hand, there
can be a bead that is jiggling between the up and down positions on
a wire. But although topological quantum computers would rely on
changing the shapes of the wires, the shapes themselves are
definite. They cannot jiggle themselves into loops that are
simultaneously open and closed, in what you might call “sort of
loops.” A wire is either a loop or it isn’t.
The interesting thing to notice is what happens if you
superimpose all the versions of a quantum abacus from the various
realities it’s in. The superposition changes depending on the role that
topology plays in your computer. The complexity of a nontopological
(nonknotted) superposition is linear, which means simply that it’s
exactly as complicated as you’d expect from summing up the
complexity of its parts. If, however, you superimpose knotted,
topologically interesting abacuses, the result will not be linear,
because not all the intermediate states are possible. If you
superimpose a loop and a nonloop, you don’t get a distribution of
possibilities including a “sort of loop.”
Now recall that we just mentioned something else that isn’t linear:
relativity. Space-time is curved, not linear. This is what we observed
when criticizing the prediction of the Unruh effect. The quantum state
of gravity, which can be thought of as the superposition of the
quantum abacuses that either are or describe (depending on your
philosophical preference) the state of space-time and/or gravity in
the universe, is not linear. Could that be because gravity’s abacuses
are knotted?
So here’s a scenario of what might have happened in the history
of our universe. These days we take matter for granted as a
dominant component of reality, but when the universe was younger,
it was too hot for matter to be commonplace. Eventually the universe
cooled down and matter became an important phenomenon. With
matter came chemistry, and with chemistry came life and evolution,
leading to smart aliens who experienced their own Moore’s laws,
which drove them to develop gravitonic quantum computers.
As it happens, the amount of dark energy in the universe reduced
rapidly just as matter became important. In terms of our thought
experiment, that was not a coincidence. It was because the aliens
used dark energy as a resource to run their ultimate computers in
much the same way we devour oil to run our cars and jets.
Before we continue, there is an even more daunting issue that
sneaks up on anyone who tries to explain away the cosmological
constant problem: Why now? I had tried to tackle this when I was at
Stanford. The surprising thing is that today the cosmological
constant/dark energy is not zero but proportional to the dense
matter; it seems to be tracking the amount of matter. The approach
Lanier and I took in our thought experiment was to turn the question
of “Why now?” into a new form: Why here? Recently a consistent
cosmology with no dark energy has been proposed by a large
number of respected cosmologists, like Joe Silk and Subir Sarkar,
who argue that if we live in a region void of the excess dark matter
then we can do away with dark energy. Likewise, our aliens come
from a region void of dark energy because this void represents a
biosphere of computational activity. According to a detailed multidata
analysis of the latest measurement of the cosmic microwave
background, the Sloan Digital Sky Survey and type Ia supernovae
observations, such a region exists. Working with my colleagues at
Penn State and CERN, we were able to show that all that data is
consistent with a region void of dark energy spanning some two
hundred megaparsec. This is a huge region; for comparison, ten
kiloparsec is about the radius of a typical spiral galaxy like our own
Milky Way, and a parsec itself is already pretty big, at more than
three light-years. So we at least have a spot where these aliens
could live with their computers. Everywhere outside the aliens’
existence, where there are no computers, the dark energy would
exist.
How would this work? In our idea, aliens use the vacuum state as
a “reservoir” of qbits. As they do computation, they tie more and
more complicated knots in the gravitational, or space-time, quantum
state of the universe, which we can think of as all those knotted,
superimposed abacuses. This has the effect of “using up” the
curvature due to the vacuum energy.
Thus, instead of seeing the enormous value we would predict
naively from thinking about a vacuum state that hadn’t been
tampered with, we instead see the tiny cosmological constant that
Saul Perlmutter measured. If this is correct, then that would mean
that the aliens have almost, but not quite, maxed out the
computational potential of the universe!
Our thought experiment didn’t quite lead us to an idea of how
alien gravitonic computers could work, but we do have some hints.
Vacuum energy universally interacts with the gravitational field.
There are some important direct channels through which this
interaction between gravity and the vacuum energy takes place. It
was shown long ago by Einstein and Élie Cartan that fermionic
matter (like electrons) will universally mediate vacuum energy with
gravity. But there are many types of fermions, of course, and
theorists have shown that chargeless fermions (similar to neutrinos)
will have the most resonance with the vacuum energy. To highlight
the computational mechanism, consider an analogy with superfluid
helium. Helium atoms behave like inert fermions, and at very low
temperatures a quantum interaction between the helium atoms make
them condense into a superfluid with astonishing emergent
properties; one is that the superfluid has a negative pressure, just
like we considered before for vacuum energy. A quantum field theory
calculation successfully predicts this about helium, and Nobel Prizes
were earned for the experimental verification of this superfluid state.
It turns out that coincidentally gravity mediates exactly the same type
of quantum interaction between chargeless fermions!
Therefore the analogy between superfluid helium is quite relevant
and important for us because, if a region of space-time is cold
enough, the correlations between inert fermionic matter can act as a
self-organized medium that tethers the ribbonlike structure of the
gravitational-fermionic vacuum energy.1 Very intelligent aliens can
achieve quantum computation by exciting the various energy and
spin states in this fermion-gravitonic superfluid. How cold does this
fluid environment have to be? A back-of-the-envelope calculation
reveals that qbit states can be manipulated by turning vacuum
energy into thermal radiation so long as the temperature of the
superfluid is less than the temperature of the cosmic microwave
background, three degrees above absolute zero. However, the
region will have to be filled with the neutral fermionic matter that the
aliens use to fuel the process. This material can be gotten from
supernovae, which emit a huge flux of neutrinos suitable for the
purpose. Neutrinos are famously hard to catch, so the aliens will
either have to figure out how to trap and contain the neutrinos that
pass through their part of the universe, or simply have to inhabit
regions rich with supernovae.
It is plausible that the aliens would have control of hot rod
gravitational wave detectors, not unlike the laser-interferometer
gravitational-wave observatory, or LIGO, that detected gravitational
waves first in 2015, that can detect and write information into the
topological vacuum state using coherent gravitons, in the same way
that our transistors can detect electron currents and switch gates.
The gate switching would be powered by the gravitons’ states
emitted and absorbed from the vacuum. There have been other
proposals to read and write gravity waves. UC Merced physicist
Raymond Chiao has proposed a variety of curiously shaped
superconducting antennas that might be able to read and write
gravity waves.
To push weirdness to the extreme: maybe there could eventually
be some way to safely create and manipulate artificial black holes
that are analogous to anyons. Maybe moving them around carefully
could perform quantum computation. At any rate, ideas about the
actual implementation of a topological graviton computer are entirely
speculative.
One oddity of gravitonic computers is that while the aliens and the
interfaces they use to operate their computers exist locally, just like
ordinary objects or like our bodies, the computation itself is not
localized. The vacuum state is necessarily nonlocal since the qbits
are entangled across our cosmic horizon (the portion of the universe
that can be relevant to us as limited by the speed of light), and that
region spans three thousand megaparsecs. That’s one big machine
room!
It should be, because the computers are stupendous. Paola A.
Zizzi of the Università di Padova has calculated that universal
information capacity in vacuum energy is approximately 10,120 bits.
That’s a lot of bits. How many qbits of computation are required by
alien virtual worlds? If we guess how many civilizations there are, we
can then estimate the average size of a planetary gravitonic
computer. If we conservatively assume just one gravitonic computer
per galaxy, we end up with googol-scale capacity—10,100 qbits—for
an average alien gravitonic computer. When we try to fantasize what
seems to us to be the absolutely ultimate planetary VR computer,
assuming that Moore’s law will run into the twenty-second century,
we still only come up with a need for a capacity that is perhaps sixty
orders of magnitude smaller.
But it’s also true that you can never have enough computation! In
fact, there’s a reason alien computers might need to be so huge, and
it has to do with heat. Gravitonic computers give off heat just like
other computers. Where does that heat go? It is interesting that there
are a vast number of highly energetic cosmic ray events and their
source remains a mystery. Even so, there isn’t anywhere near as
much heat as might be expected from computers of the stupendous
size hypothesized here.
You might just take this as a sign that the computers don’t exist,
but let’s keep working with the idea that they do. If you want to
reduce the heat a computer generates and you have a huge amount
of memory, you have an amazing design option, which is called a
reversible computer. That means that you change each bit in the
whole computer only once, and then move on to another bit. That
results in a total record of all computation—and that’s why it’s
possible to run the computer in reverse: nothing has been lost. If you
move each bead in an extremely capacious abacus only once, you
don’t generate the heat you would by moving each bead repeatedly.
You can think of it as saving all the information in a tidy way instead
of dispersing it. This is also a nice example of Claude Shannon’s
famous principle that information and entropy are related.
What will become of our cosmic khipu weavers? They might end
up like Maxwell’s demon, a thought experiment important to the
development of the second law of thermodynamics that remains
important precisely because Maxwell’s demon can never exist. Or
maybe the weavers do exist and we’ll meet them. Maybe we’ll
eventually weave our own gravitonic khipu.
Of course it’s possible the aliens don’t exist but that graviton
weaving will still turn out to be part of the solution to the mystery of
dark energy. But that would suggest a natural, self-propelling
process that relies on information storage—and that sounds like a
definition of something very much like life.
14

INTO THE COSMIC MATRIX

All living things are born and change throughout their lifetime. The
expanding universe is like this. Because the universe’s space is
expanding in time, when we reverse the clock, it returns to an epoch
where time itself was born. For many years, I tried to get my mind
around the question: What can exist if time ceases to exist?
To delve deeper into this and other issues about the early
universe, it’s useful to address a few common misconceptions about
the big bang itself. Perhaps the biggest misunderstanding is that the
big bang was some cataclysmic explosion that fueled the universe’s
expansion, where galaxies went flinging away from each other. In the
standard big bang theory predicted by general relativity, the universe
is assumed to be filled with a hot and dense gas of matter and
radiation, whose energy and pressure source the expansion of
space-time itself. We’ve discussed before that general relativity
provides an expanding solution to its equations; called the FRW
model, for its discoverers Alexander Friedmann, Howard Robertson,
and Arthur Walker, it was for a long time the standard model of
cosmology. According to that solution to the equations of general
relativity, the expansion of space is not an “explosion” but actually
gradual in time. Even more interesting is that time ends in a
singularity of infinite curvature and energy density. What is this
singularity really trying to tell us?
To try to answer the above question, Stephen Hawking and Roger
Penrose provided a theorem that the FRW expanding solution will
always suffer a singularity. At the singularity, the curvature goes to
infinity, marking a breakdown of general relativity as a valid
description of the space-time structure. The Hawking-Penrose
theorem was based on a powerful equation discovered
simultaneously by Indian theorist Amal Kumar Raychaudhuri and
Soviet theorist Lev Landau. These Landau-Raychaudhuri equations
relate paths of observers in a curved space-time, called geodesics,
to singularities. Hawking and Penrose implemented these equations
to prove that geodesics in an expanding space-time will exhibit
pathologies as they approach the infinite curvature singularity at the
beginning of the big bang. A useful set of geodesics to use as a
diagnostic for singularities are light rays. In a flat space-time two light
rays will follow parallel lines and never cross each other. In a curved
space-time the paths of light rays can twist, focus, or even diverge,
depending on the warping of the space-time. Hawking and Penrose
showed that as light rays approach the earliest times toward the
infinite past in an expanding space-time, their geodesics terminate.
This geodesic incompleteness signals the infinite curvature and a
singularity as time goes to zero.
A similar story happens for black hole singularities with
geodesics, and many interpret singularities as a breakdown in the
validity of general relativity at the singularity. Of course, all theorems,
including the Hawking-Penrose theorem, do come with assumptions,
or axioms, and axioms aren’t necessarily true. Sometimes they can
be relaxed; you may know that the parallel postulate in Euclidean
geometry can be relaxed, and rather than breaking geometry,
actually points the way to two other kinds, hyperbolic and elliptic.
Perhaps if we can relax one of them this could be a way out of the
inevitability of the cosmic singularity. One assumption of the
Hawking-Penrose theorem is that matter is classical. Keep in mind,
though, that the classical matter in the world is made up from
quantum matter. And cosmic inflation uses quantum matter and the
inflation field, and it may help resolve the singularity. But even the
quantum state of inflation does not rescue the universe from the
singularity.1 The onset of inflation still remains unresolved. Over the
last four decades cosmologists have been investigating various
mechanisms based on new physics that could alleviate the big bang
singularity. There are a handful of promising approaches my
colleagues and I have pursued.
Aside from potentially satisfying the urge to understand the birth
of the universe, pursuing the problems facing big bang physics
provides an opportunity for physicists to guide themselves toward a
theory of quantum gravity and to test the validity of those they find.
At one level we have constructed theories of quantum gravity by
attempting to make the principles of general relativity and quantum
mechanics consistent with each other. But the issues that the early
universe presents can serve as guideposts to a more fundamental
theory—and it may not even be quantum gravity. As we go back in
time, the universe approaches a length scale called the Planck
length, which is thirty-five orders of magnitude smaller than a meter.
Here, we expect gravity and quantum mechanics to speak to each
other. In some quantum theories of gravity, space-time emerges from
a pregeometric phase, a vision of reality that would force us to
conceptualize matters where there is no space-time to refer to. In
other models, the expanding universe must arise from primordial
atoms of space-time. In those cases how these “atoms” interact to
give us an expanding universe will be the prize hunt. As we
explored, both string theory and loop quantum gravity have
ingredients to avoid certain space-time singularities, and models of
the early universe have been proposed and developed to transcend
the singularity.
Understanding the evolution of the earliest universe provides the
foundation for a sequence of important events for other forms of
evolution to take place, such as star formation, which is central to life
as we know it. Therefore, understanding the earliest stages and
ideally the origin of the universe may give us new insights into the
universe we currently inhabit. Just like the genome reveals new
secrets about an organism, could it be that the pre–big bang
universe can shed light on our universe in its current epoch?
Recall that a successful theory of the early universe must solve a
handful of cosmological problems that currently have no confirmed
solution. A suite of observations have made it clear what these
conundrums are. In addition to the big bang singularity problem,
these are the horizon and flatness problems (which João Magueijo,
among others we’ve met, are working on), the fine-tuning problem,
as well as the question of the origin of large-scale structure such as
galaxies and the CMB. We have seen how cosmic inflation, without
departing too much from the principles of general relativity and
quantum field theory, has already been able to solve some of these
problems, providing answers to the horizon and flatness problems,
as well as giving an explanation for the universe’s large-scale
structure. You may wonder, if cosmic inflation does such a great job
of explaining our current universe, then why seek alternatives? One
answer is that it can’t explain the singularity and fine-tuning
problems. Another is that, even if an alternative theory turns out to
be wrong or simply less successful than the explanations we already
have, exploring it can still give us a new perspective that can
improve the more conventional theory. What’s more, I believe that
the singularity problem that plagues inflation points to the source of
all the cosmological problems—if we can resolve singularities, we
might resolve it all.
Over the years my colleagues and I tried to alleviate some
conceptual and mathematical problems that plagued cosmic inflation
by constructing a superstring theory inflationary model, and we kept
hitting roadblocks. It was only when we researched alternatives to
inflation within the context of superstring theory that we got new
insights into moving forward with the inflationary roadblocks—an
example of the value of gaining an outsider’s perspective. The fact
that the universe is expanding actually limits the options as to what
may have occurred before it started to expand. This boils down to
the question of whether the universe had a beginning or is eternal.
We will now venture into discussing promising theories for
understanding how the earliest stages of the universe could have
emerged. Keep in mind that the theories that we will consider may or
may not need inflation. And as we venture to get a sense of the
future of the physics of the early universe, we should take stock of
the two general and divergent hypotheses behind approaching a
theory of quantum gravity.
Hypothesis 1: We should quantize gravity the same way that
proved to be successful in other systems, such as quantum
electrodynamics and quantum chromodynamics. In this approach,
one identifies the classical system, such as the state space (or
phase space) and applies rules for quantization. For example, in
particle mechanics the phase space will comprise all positions and
momenta of the dynamical particles. These measures commute with
each other, which means that the order in which we multiply with
them doesn’t change the outcome.2 Quantization rules impose that
position and momenta don’t commute, however, and the classical
phase space gets promoted to a Hilbert space. In Hilbert space, the
central objects are not finite vectors (of the kind that described
momentum in our discussion of phase space early in the book), but
rather infinite dimensional vectors. These are otherwise known as
wave functions and correspond to a probability distribution.
Hypothesis 2: General relativity is a classical theory that should
not be directly quantized according to hypothesis 1. Instead, at
shorter distance scales there are more fundamental quantum
degrees of freedom that give general relativity as a long-distance,
low-energy classical theory: this is the quantum emergent principle
at work. This is similar to how fluid dynamics emerges as a long-
range theory of interacting atoms. As we discussed, string theory
does not quantize general relativity, but general relativity instead
emerges as a low-energy effective theory (that is, a theory that
describes effects but not causes of an observed phenomenon—in
this case, gravity), albeit in ten dimensions.
We’ve explored both so far in this book. But let’s turn now to a
promising approach to the early universe that pursues new
directions. One of the interesting things about this approach is that it
addresses primarily the big bang singularity and the other
cosmological problems that cosmic inflation fails to address. This
isn’t unique to cosmic inflation: We will see that all approaches are
limited and require concepts and tools that go beyond the current
theories that we are considering.
In a pioneering publication in 1989, two theoretical physicists, Robert
Brandenberger and Cumrun Vafa, presented a new approach to the
early universe by attempting to solve three questions in one fell
swoop. The key insight into the BV mechanism, as their result came
to be called, harkens back to the concept of duality in quantum
mechanics. They not only questioned the initial spatial singularity
and infinite temperature at the bang but also related the solutions to
those problems with something we take for granted: the
dimensionality of space. The basic laws as we know them would be
different if space were not three-dimensional. Electric and
gravitational fields would not fall off inversely proportional to the
square of the distance, which would affect all the chemistry
necessary for the world as we know it. If we lived in two spatial
dimensions, for example, life would not exist as we know it; for the
function of carbon-based life depends on three-dimensional structure
of folded proteins. In the spirit of relativity, we can think that the
dimensionality of space-time is not absolute and treat the fact that
we live in three dimensions as a physical condition that emerged in
the early universe.
Of course we’ve seen that string theory is a quantum theory of
gravity that requires nine spatial dimensions in order to be quantum
mechanically consistent. What Brandenberger and Vafa realized was
that string theory had the correct ingredients to address the other
cosmological problems in one fell swoop by first asking how the
unique properties and symmetries of quantum strings address these
cosmological conundrums. Consistent with superstring theory—a
theory that fuses supersymmetry with string theory—Brandenberger
and Vafa considered an early universe in nine spatial dimensions.
Instead of the early universe being occupied with a thermal state of
particles and radiation, it is filled with a thermal state of strings.
Because strings are extended objects, and unlike particles, they
have the special property that they can wind around a spatial
direction. So, string theory has both oscillatory strings, and wound-
up strings called winding strings. To get a feel for the differences
between those two types of strings versus particles, consider the
geometry of a two-dimensional torus, which looks like a doughnut. A
particle can only move along the surface of the doughnut, but a
string can both move along the surface of the torus and wrap around
one of the cycles of the torus. These winding states, like rubber
bands, have tension energy; if they are left to their devices they will
cause the torus to collapse. This is also true if you are speaking
about the kind of hypertorus you’d find in a nine-dimensional world
such as Brandenberger and Vafa were considering. But according to
general relativity the radiation that lives in space makes space
expand. So, in a stringy universe there exist three types of matter:
winding strings, string loops, and radiation. In such a universe space
will expand and the winding energy will force the space to collapse
after it expands too much.
Under ordinary circumstances, according to general relativity,
such a universe will not be able to avoid a singularity. But string
theory has a new ingredient that tells a different story: a symmetry
we previously discussed (I used it in my thesis) called target space
duality, or T-duality. This symmetry states that physics in the smallest
possible region of space is equivalent to physics in the largest
possible region of space. This happens because string theory has
both matter and winding configurations present. If we perform a
transformation that swaps the roles of these two states, it is
equivalent to swapping motion in a large region of space and a small
region of space. For example, if the torus universe has a radius R
then the transformation will take R to 1/R; nevertheless, string theory
remains unchanged. If the radius is cosmologically large, then
strings moving will experience the same physics as if they were in a
small space 1/R.
FIGURE 26: This diagram shows how the BV mechanism avoids the
singularity as the radius goes to zero. We enter the T-dual phase instead
of diverging with infinite curvature.

Here is an intuitive way to understand this. Let’s imagine that the


universe shrinks to distance scales that are microscopically small,
such as the string length. In the standard big bang picture, because
only particles were present, the temperature would tend to infinity as
the universe got smaller—exactly the kind of singularity we’d like to
avoid. But in a stringy universe as the temperature gets high enough
as the scale shrinks, the energy of particles starts getting fed into the
string oscillations. Eventually all these oscillation states get occupied
and the universe approaches not an infinite temperature but a
maximum possible temperature, known as the Hagedorn
temperature. At the same time as we approached the R=0
singularity, there is a dual description of the universe in which the
physics is the same as having winding modes moving in a large
nonsingular universe. In a nutshell, the Brandenberger-Vafa
mechanism prevents the universe from reaching the singularity by
having a symmetry in the physics between winding string modes and
particle modes and swapping small with large distances. If the
universe can never become infinitely large, then by symmetry it can
never go to zero length and will avoid the spatial singularity found in
standard big bang cosmology.
But by using the same physics of winding states, the BV
mechanism goes beyond this by predicting why we live in three
spatial dimensions. Again, the argument uses the properties special
to strings. Understanding why we live in three dimensions is as
simple as understanding the question: In what dimension are two-
point particles most likely to collide? In three dimensions if we set
two particles (particles have zero dimension; strings have one
dimension) off in a random direction, they are less likely to run into
each other than in two dimensions. But in one dimension then the
particles are guaranteed to run into each other.3 A similar argument
applies to strings, which are one-dimensional objects. It turns out
that in three dimensions strings are most likely to collide with each
other without the collisions being unavoidable; in anything larger
than three spatial dimensions, strings are likely to never meet. In the
Brandenberger-Vafa model there are also strings that wind in
opposite directions relative to each other. If these strings run into
each other they, like particles and antiparticles, will annihilate into
radiation. Consider a three-dimensional torus of both winding strings,
which keep the space from expanding due to their tension exerted
on the space. If these winding modes annihilate then the space will
expand according to general relativity. In a ten-dimensional universe,
Brandenberger and Vafa found that the strings in the extra six
dimensions never get to annihilate. As a result, three dimensions
become our observable expanding universe while the other six
remain wound up by the winding modes.
So now the BV mechanism tells us that our universe is a ten-
dimensional, string-dominated space-time with all dimensions
starting out microscopic. Applying the principle of maximal symmetry,
at the beginning all space-time dimensions and string states are on
equal footing.4 Therefore, strings and antistrings will wind around all
nine spatial dimensions. These dimensions will remain tiny until
winding strings annihilate each other in three spatial dimensions.
This will cause a large three-dimensional space to expand. Because
of T-duality and the finite Hagedorn temperature, there is no big
bang singularity. This suggests the universe did not emerge from a
big bang singularity (that is, from a situation where the radius was
zero and then began to increase).

FIGURE 27: A schematic representation of a dimension being


compactified by winding strings.

The BV mechanism also provides a way of seeding the large-


scale structure in the universe that is different from the way that
inflation describes the process. In inflation cosmology it is quantum
vacuum fluctuations that initiate cosmic structure. In the BV
mechanism it is the thermal waves generated by the undulating gas
of strings that give the observed nearly scale-invariant fluctuations
seen in the CMB. Both mechanisms generate a spectrum of
gravitational ripples of space-time that leave an imprint on the
polarization of light in the CMB. These gravitational waves are
predicted to leave in their wake a pattern on the CMB photons by
creating an overall curling, like a pinwheel, of the light’s polarization,
called B-mode. Observational cosmologists have been on a hunt to
find this faint signal, pushing the envelope of the most advanced
detection technology known. Currently the Simons Array telescope
led by my colleague and friend Brian Keating is being built to finally
detect the primordial gravitational waves. But although both models
predict the waves, they don’t predict the same pattern in the B-mode
polarization; observing the pattern could tell us if our universe
underwent inflation or something more akin to BV. The predictions of
gravitational waves between inflation and BV differ in that the shape
of the power spectrum tilts in the opposite direction. In inflation the
power spectrum is said to be red. This simply means that the longer
wavelength perturbations have slightly less power than the shorter
wavelength ones. BV has an opposite blue power spectrum.
The BV mechanism elegantly paves a way to solving another big
problem in string theory that I discussed in a previous chapter on
quantum gravity. Recall that the information about the extra
dimensions in string theory show up in our four-dimensional world as
a large number of fields that could have disastrous effects in our
world, called moduli. If string theory is correct, it must provide a way
to freeze, or stabilize, these moduli so that they do not disrupt the
observations of the CMB and large-scale structure formation.
My colleague Subodh Patil and his collaborators elegantly
discovered an intrinsically stringy mechanism such that the string
modes generate a force to stabilize all the moduli in string theory,
which made use of new symmetries that are unique to string theory.
Patil is a bassist, and I am curious if he imagined the moduli
stabilization mechanism as analogous to how a bass line can serve
as an anchor for a melodic line. The mechanism also addresses a
worry that I always had about the BV mechanism, one that I
remained silent about for years. Superstring theory relies on
supersymmetry; a symmetry puts fermions and bosons on the same
footing, for its consistency. However, our world is not
supersymmetric since matter is distinct from its force carriers. In
particular, a time-dependent space-time such as our expanding
universe is not consistent with supersymmetry. So, trusting the
validity of these intrinsically supersymmetric string modes within a
cosmological background seems a bit ad hoc. However, Patil and
collaborators showed that it was exactly the breaking of
supersymmetry that provided the forces for the stabilization of
moduli.
The BV mechanism provides a compelling framework that
successfully goes beyond the standard big bang theory. Some
unresolved issues persist because the BV model assumes the string
universe will exist in the Hagedorn phase for an indeterminate
amount of time, then at some point in time the winding modes
annihilate, and the universe begins to expand from a nonsingular
state. How long does the universe remain in the high-energy stringy
state? What placed the pre-expansion universe in such a state? The
question of initial conditions persists in this version of the BV
mechanism. One possible way out is to invoke a cyclic cosmology.5
In this framework the universe undergoes a series of expansions and
contractions long enough for the winding modes in three spatial
dimensions to annihilate, and our three-dimensional universe
expands and gets macroscopically larger than the other six.
Current technical and conceptual details are being researched
that also touch on other profound questions about the early universe.
One big issue that goes beyond BV and other approaches to the
early universe concerns the emergence of space itself. In fact this
subject is currently the state of the art in quantum gravity research.
There are a handful of promising approaches to emergent space-
time but I will restrict our discussion to an avenue that extends the
BV approach naturally, while also illustrating how to improve a
theoretical model. There currently exists no complete theory from
which we can obtain an expanding cosmology from a pregeometric
state, so I will argue that a promising direction comes from the
general framework of noncommutative geometry. But what would we
learn about cosmology from a space-time that emerges from such a
pregeometric phase?

In 1989 my friend the late renowned theoretical physicist Joe


Polchinski realized that strings were not the only fundamental
objects in string theory. He did this by solving a long-standing
problem about how to apply T-duality to a version of strings that have
no windings. These are called open strings. Polchinski discovered
new extended objects that are membrane-like, which he crowned
Dirichlet-branes (or D-branes), that open strings can collectively end
on.6 These D-branes come in many dimensions; for example, a 2-
brane is a two-dimensional hypersurface otherwise called a
membrane. For part of my PhD dissertation work, I, along with
Damien Easson and Robert Brandenberger, showed that the BV
mechanism works even when we include a spectrum of D-brane
states in string theory.7 When I gave my thesis defense, one of my
examiners was David Lowe. Lowe is a string theorist, and he asked
why we did not include D0-branes. This is a zero-dimensional object
that strings attach to. I did not have a good answer, except that D0-
branes do not have winding modes, like their higher-dimensional
cousins. But I’ve reconsidered my response. It turns out that going
down that path, incorporating D0-branes to reimagine the early
universe, reveals an interesting new possibility—that the universe
could have existed without reference to space itself.
The theory that describes D0-branes falls into a class of quantum
geometric theories called noncommutative geometry. What is
intriguing is that a handful of approaches to quantum gravity all have
some semblance to a pre-space, where geometry itself is fuzzy, or
noncommutative.
FIGURE 28: A membrane is classically unstable because it is more
energetically favorable for infinite spikes to form on the surface than to
keep the membrane surface smooth. This instability posed problems in
quantizing the membrane as a fundamental object in quantum gravity.

In ordinary quantum mechanics the spin of the electron in the x


and y direction cannot be known simultaneously. For example, the
more certain you are about spin x the less certain you are about spin
y. Suppose the determination of different directions in space were
noncommuting. Then the usual notion of a smooth, continuous
space-time is no longer valid. Imagine a noncommutative space
where a location is restricted to be at a unique point in the x
direction, then a definite location in the y direction will be uncertain.
In general, a location in space becomes fuzzy in a noncommutative
space. There are a handful of proposals in formulating
noncommutative space-times. The theory of D0-branes is known as
matrix quantum mechanics.
In matrix quantum mechanics, there are nine matrices, each
representing the dynamics of the D0-branes. The D0-branes have a
potential energy made up of these interacting matrices that do not
commute with each other, which encodes the uncertainty in the
spatial dimensions. What I find most interesting about matrix theory
is that many formulations of quantum gravity point to the same
theory. This coincidence of a quantum theory with no a priori spatial
dimensions is a fertile ground to investigate a pre–big bang phase
where our smooth, classical, and expanding universe may have
emerged from a reality in which space itself did not exist. At this
impasse, I leave this speculation of a working emergent
cosmological framework as a future project perhaps successfully
pursued by a brilliant young cosmologist.
15

THE COSMIC MIND AND


QUANTUM COSMOLOGY

Several years ago, a handful of scientists and I received an


invitation to have a discourse with Dr. Deepak Chopra at the
Carnegie Institution. The other physicist attending was Nobel
laureate Frank Wilczek, so I couldn’t resist the chance to be in great
company. However, I approached the interview with the fear of being
further sequestered from the community of scientists—many who
held Chopra to mix mysticism with physics. Chopra and his
colleagues have been proponents of the idea that consciousness
creates the physical universe.1 It is well known that certain scientists
have debated with Chopra and even criticized his ideas as “out
there.”
Chopra is a brilliant medical doctor and effective communicator of
science to the masses, and I was silently impressed and inspired by
his willingness to debate unconventional ideas with renowned
scientists. But he was clearly an outsider to the enterprise of
research scientists. I was sympathetic to the resistance he received
because he was not a member of the club, made claims that could
not—yet—be experimentally supported, and was considered to be a
charlatan by a handful of vocal, skeptical voices. And I always
enjoyed watching Chopra work the crowd and effectively debate with
other scientists. Walking into my interview, I knew that he would
certainly ask me about consciousness, quantum physics, and a topic
that he has even published on recently with others, called cosmic
consciousness.
In the middle of my interview, Chopra dropped the “C-bomb” on
me. With his resonant baritone voice he asked, “So, Stephon, do you
think the big bang came from consciousness?”
Like a skilled coward, I dodged his question. I said, “I’ll take off
my scientist hat and put on my Stephon-as-a-person hat and say
that…” What I wanted to say was: “Deepak, I got into physics
because I wanted to understand how this basic, direct experience of
consciousness was connected to the fabric of the universe.” Instead,
I gave a weak response and said, “I think that a future generation of
brilliant physicists should be brave enough to tackle the question.” I
chickened out and Chopra knew it.
I wasn’t alone. At that time I was unaware that a handful of the
architects of quantum mechanics, including Bohr, Heisenberg,
Schrödinger, Wigner, and von Neumann, had been influenced by the
question of consciousness in the physical world. Schrödinger was
especially influenced by the work of German philosopher Arthur
Schopenhauer and the Vedas, which posited the existence of a
universal mind that contains all individual minds and the physical
universe. Combining those with his own work, Schrödinger imagined
the quantum wave function to be part of an undivided cosmic whole.
We have explored the consequences for our universe of the
quantum, invariance, and emergence principles. The laws that follow
from these principles, general relativity and quantum field theory,
precisely predict a universe that expands from the big bang into a
seemingly structureless early universe that vibrates with sound
waves of energy and ultimately grows into the web of stars, planets,
and galaxies. We don’t have a complete understanding of the bang,
the cause of the waves, and the emergence of space-time itself. And
as we’ve seen, if we want to transcend the big bang singularity, we
may very well need ideas that go beyond our current principles.
And those might not even be the biggest issues we face. The
expansion of the early universe linked with the flow of entropy
necessary for biological life is a hint at a deeper interdependence
between life and the quantum universe. Did life emerge in the
cosmos through a series of accidental historical events? Is there a
deeper principle beyond natural selection at work that is encoded in
the structure of physical law? And on top of that, the question that
bothered Schrödinger and that got me into science in the first place:
What is the relationship between consciousness and the fabric of the
universe?
I am fully aware that I risk being written off as an oddball crank by
the positivists in the room, because answering these questions might
call into question the idea that the world out there is independent of
us being there. But I must eat my own words to Deepak Chopra and
embrace the stigma of weaving together theoretical ideas to
entertain that question. In this final chapter, I am going to engage in
an exercise of theoretical dreaming—into the principle of blackness. I
am now going to pursue a speculation that intimately relates the big
bang to the most complex entities that emerged from the universe:
human beings endowed with conscious awareness.
This begins with the elephant in the room, the measurement
problem in quantum mechanics. Quantum systems exist in a
superposition of states until a measurement collapses the wave
function into one unique state. In his book Mathematical Foundations
of Quantum Mechanics, John von Neumann, the father of the
modern computer, proved that when a quantum system exists in a
superposition of states, a chain of measurements ultimately leading
to the consciousness of an observer is what collapses the wave
function into one definite state. He argued that this collapse by
consciousness cannot be consistent with the mathematical
framework of quantum mechanics, especially the linearity of
quantum mechanics. This interaction between consciousness and
quantum mechanics has to happen outside the constructs of
quantum theory itself—unless consciousness is part of quantum
physics from the start. We’ll get back to this.
As we have discussed, in his book What Is Life? Schrödinger
makes three key observations about what differentiates living from
nonliving matter subject to the laws of physics including quantum
mechanics. First, he predicted the basic helical structure for DNA
using ideas from quantum mechanics that describe the periodic
lattices found in metals. Second, he argued that living things fight
against entropy, otherwise known as negentropy. Both predictions
inspired the next generation of biochemists and biophysicists and
continue to be foundational. And third, Schrödinger speculates about
what it means that some (if not all) biological life has consciousness.
He asks what is at first glance a strange question: Why are there
many minds, each having their unique conscious experiences?
When Schrödinger was writing this work eighty years ago, he
recognized that his era lacked a scientific account of consciousness,
and so he resorted to philosophical and metaphysical ideas from
Arthur Schopenhauer and Vedic philosophy. Both sources believe
that the universe and all that occupies it carries varying forms of
awareness. This view is often known as panpsychism. Panpsychism
posits that consciousness is an intrinsic property of matter, the same
way that mass, charge, and spin are intrinsic to an electron. So
according to panpsychism, the electron and all substance come
equipped with their own internal protoexperience of being an
electron. This might sound crazy. Definitely there’s a question about
how an entity, say an electron, can have its own internal experience
without resorting to an electron brain. The answer requires new
physics or a fresh perspective on known physics. And we find a clue
from African philosophy.
While there was an absence of African philosophy in my formal
education, my musical colleague, legendary bassist Melvin Gibbs,
introduced me to a view of the creation of the universe presented by
the Bantu-Kongo people of West Africa that predates our modern big
bang theory—but it has more, including an extra conceptual key to
help us understand how to relate the quantum, cosmos, and
consciousness with one another. In the Bantu cosmology the
universe started in a state of nothingness called mbungi. Here
nothingness includes the absence of space and time. Physical
objects, such as particles and fields, usually exist in space-time. So
mbungi is a prephysical state that is divided into what manifests as
the physical, spatiotemporal world and a universal consciousness.2
In the state of mbungi both the physical and conscious awareness
are complementary, and have close semblance to the yin and yang
in Taoism. Therefore, mbungi finds a natural home in quantum
complementary in the context of the entire universe. Translating this
into the language of cosmological physics, we can define a
pregeometric universe as a quantum state that contains the
potentiality of space-time and a fundamental form of consciousness
as complementary pairs. To make a cosmological complementarity
concrete we will need to actually have a formalism for quantum
cosmology—a wave function of the universe.
In 1985 Stephen Hawking and James Hartle published a paper
entitled “Wave Function of the Universe,” which implemented the
Schrödinger equation associated with quantum gravity known as the
Wheeler-DeWitt equation. Unlike the original Schrödinger equation,
which gives the time evolution of the wave function, the Wheeler-
DeWitt equation is timeless.3 Hartle and Hawking found a wave
function of the universe, which is now famously called the Hartle-
Hawking state.
The wave function of the universe is not some theoretical
playground; it actually corresponds to the quantum state that
underlies cosmic inflation that we discussed in an earlier chapter. In
fact it was shown by cosmologist Alexander Vilenkin that the wave
function of the universe can undergo a process of quantum tunneling
from a state of nothingness—where space vanishes—to an inflating
space-time.4 Recall that a quantum system can go through barriers
that are forbidden by classical physics. In this case the quantum
universe can tunnel from a state of no-space, which is inaccessible
to classical physics, into an inflating space-time. Because the Hartle-
Hawking-Vilenkin state is connected to inflation, it is taken quite
seriously by the cosmology community as a benchmark for doing
calculations that correspond to satellite observations. Despite its
pragmatic importance, the Hartle-Hawking-Vilenkin wave function
presents a deep conceptual problem about the nature of time and
the emergence of space at the big bang. Near the end of his life
Hawking stated, “Asking what came before the Big Bang is
meaningless… because there is no notion of time available to refer
to.… It would be like asking what lies south of the South Pole.”
Renowned Stanford University quantum cosmologist and co-
inventor of inflation Andrei Linde gives us a hint of how to resolve the
problem of understanding how the universe emerged from a state of
no-space. Linde focuses on the fact that the Wheeler-DeWitt
equation of quantum cosmology is timeless, which implies that the
universe is “dead.” Linde proposes the way out of this conundrum is
to link consciousness with space-time. In a remarkable article Linde
asks, “We cannot rule out the possibility a priori that carefully
avoiding the concept of consciousness in quantum cosmology
constitutes an artificial narrowing of one’s outlook.… Is it not possible
that consciousness, like space-time, has its own intrinsic degrees of
freedom, and that neglecting these will lead to a description of the
universe that is fundamentally incomplete?”5
But given this hint, how does the universe as we know it get
realized at the big bang? This question is still up for debate, and I will
argue that the wave function of the universe undergoes self-
observation, a form of cosmic protoconsciousness, in the spirit of
how the physical world emerges from mbungi.
The question, then, is how is it that this cosmic
protoconsciousness can be timeless and spaceless? Explaining it
seems difficult: neuroscience doesn’t even have a complete
explanation for where human consciousness comes from. My
colleague David Chalmers is a leading researcher in the study of
consciousness and pioneered a concept that he called the “hard
problem of consciousness.” In a nutshell, while neuronal activities of
the brain can correlate with various perceptions and conscious
experiences, it cannot explain our private, subjective, internal
experiences of perception, self-awareness, emotions, and other
experiential states of consciousness. The point the panpsychists
make is that maybe neuroscience has set its sights too low. The
Vedic system posits that the universe comes with its own singular
consciousness, known as Brahman. With this concept, Schrödinger
was able to answer his question about the multiplicity of individual
minds if there was one universal mind by imagining that the
multiplicity of individual minds are actually a hall of mirrors reflecting
the one universal mind. For years, I was fascinated by this claim but
didn’t understand how many minds could be equivalent to one. And
Schrödinger resisted making any further physical connections to his
conviction about minds in the universe. What he failed to realize is
that the fundamental principle of quantum mechanics, superposition,
can hold the key to putting physical meat on the bone of his
conviction. At the end of the day, fourteen billion years of cosmic
evolution results in beings like you and me, endowed with the
faculties of perception and consciousness. Is the question of
consciousness solely a matter of emergence of the happenings of
the brain? Or did the early universe encode the inevitability of
conscious experience for specific cosmic functions?
The iconic theorist John Archibald Wheeler asked similar
questions and believed that quantum mechanics plays a pivotal role
in relating conscious observers to the very operation of the universe.
While it may seem absurd, this type of reasoning led Wheeler to
formulate a more bizarre version of the double slit experiment that he
called the delayed choice experiment. Maybe the wildest thing about
the delayed choice experiment is that it has been experimentally
observed.6 Here is an easy way to understand the experiment. Let’s
revisit the double slit experiment. We know that observing the
electron before it enters the slit results in particle behavior, and not
observing causes the electrons to behave like waves, due to
interference. But what if we take the observation device away from
the slits and place it at the screen? With precision technology,
experimenters were able to wait till the last moment before the
electrons hit the screen to watch it. In this case the electrons went
through the slit like waves, and right before hitting the screen they
collapsed like particles—the pattern was the same as when the
double slit experiment measures the electron at the slit, which is not
a wavelike pattern, even though you might expect that we’d get a
wavelike distribution. From this bizarre behavior it means that the
electron retrocausally went back in time and went through the slits
like a particle. In other words, observers can delay their
measurement of the electron’s particle or wavelike behavior before it
hits the screen.
In his own words, Wheeler asked, “So what does the quantum
have to do with the universe? Perhaps everything, because in any
fundamental theorem of existence, the large and the small cannot be
separated.” Here we get a sense of what it means for the universe to
have a wave function. Like a quantum particle that traverses many
paths from beginning to end, our quantum universe traces out many
histories simultaneously from the bang to the present. In his idea that
he called the participatory universe, which is a delayed choice
experiment for the universe, Wheeler proposed that when conscious
observers make quantum measurements of the early universe, we
collapse the universe’s quantum wave function to a history
consistent with our existence.
As a younger, wide-eyed grad student I came across Wheeler’s
crazy ideas, such as “it from bit” (about the relationship between
information and matter) and the absorber theory (about the direction
of time) that he coauthored with Richard Feynman. Ironically, and
similar to the experience Michael Faraday had with the idea of fields,
these ideas began as outlandish but are now the norm of theoretical
physics research, especially our quest to build a quantum computer.
Nevertheless, most physicists shy away from entering Wheeler’s
rabbit hole of the participatory universe and exploring the possibility
that seemingly insignificant specks like us can have any connection
to and influence on the cosmos.
Wheeler believed that the universe implemented Darwinian
evolution—and, as I and Salvador Almagro-Moreno argued, the
entropocentric principle to create conscious life for the universe to
observe itself. This self-measurement solves the measurement
problem in quantum cosmology in one sense, because it gives a
mechanism to collapse the universe’s wave function to its current
and future state of existence. However, the universe had to wait
fourteen billion years for the first form of biological life to come on the
scene. If there was no life before then, how could the universe
measure itself? I do not know if Wheeler was aware of Schrödinger’s
musings on consciousness in the universe, but his statement points
to the underlying physics: “So what does the quantum have to do
with the universe? Perhaps everything, because in any fundamental
theorem of existence, the large and the small cannot be separated.”
This statement points to what I believe is the key insight to
reconcile the lack of life as we know it to collapse the wave function
of the universe—a nonlocal conscious observer. This nonlocality is
complementary to locality in the same way the electron’s position is
complementary to its momentum. Recall that we discussed that the
entirety of the quantum electron relies on opposing local and
nonlocal properties; it is both a wave and a particle at the same time.
Let us assume that consciousness, like charge and quantum spin, is
fundamental and exists in all matter to varying degrees of
complexity. Therefore consciousness is a universal quantum
property that resides in all the basic fields of nature—a cosmic glue
that connects all fields as a perceiving network.
Others are pursuing similar questions. Recently, Johnjoe
McFadden and others have published neuroscience research
arguing that consciousness is carried by electromagnetic field
patterns distributed throughout the brain. Different organizational
properties of fields can carry an array of conscious experiences.
They point out each neuron in the brain can generate electric fields
around their cell membrane and these individual fields can
superpose across billions of neurons creating a complex pattern, rich
in an organizational property, discovered by neuroscientist Giulio
Tononi, called integrated information. Many prominent
neuroscientists, such as Christof Koch and Tononi, believe some
form of panpsychism is necessary to address the hard problem of
consciousness.
What these proposals don’t do is consider whether
consciousness must be limited to our bodies. Our conscious
experience is local in that our internal experience is in reference to
our localized body in space and time. We might take this for granted,
but that doesn’t mean the universe does. What would it look like to
have a nonlocal conscious experience? It could not be in reference
to a place. According to the superposition principle in quantum
mechanics, we could represent the nonlocal state of consciousness
by superposing a large number of local conscious observers. If the
universe’s quantum state is endowed with a nonlocal state of
consciousness, then according to this type of complementarity, it is
dual to a superposition of local consciousness. This would resolve
Schrödinger’s paradox—why do so many minds have their own
conscious experience? All these local minds need to be superposed
to reconstruct a unitary, nonlocal, cosmic mind, like the positions of
the electron needs to be superposed to create its field. The cosmic
mind is contained in the local minds, though hidden from our
everyday local experiences. What this means for you and me is that
our consciousness contains an aspect of the cosmic mind. In Vedic
philosophy this is referred to as the Atman, or the self.
This conclusion to some might be awe-inspiring or preposterous.
When I set out to write this book, there was no way I could have
imagined presenting this argument. If you find this line of reasoning
preposterous, it is even crazier that we came into being to even be
able to ponder these questions. Giving a shout-out to the
distinguished Indian physicist: No one ever died from theorizing!
ACKNOWLEDGMENTS

When I first entertained the idea of writing this book I was


overwhelmed with self-doubt. I want to especially thank K.C. Cole
and Maria Popova for empowering and supporting me with some
important tools and encouragement to get through writing this book. I
also want to thank Mark Gould, one of the world’s most brilliant
social theorists, for his ongoing support and collaboration on
theorizing the sociology of science with me. To my editor, the
magician TJ Kelleher—thanks for yet another amazing journey.
Thanks to Lara Heimert, Kelly Lenkevich, Sharon Kunz, Liz Wetzel,
and the Basic, Perseus, Hachette team for helping to bring this book
into reality. Thanks, Brandon Ogbunu, for your inspiration, guiding
me through the very first stages of writing, and continuing to be a
soundboard all the way through. Thank you Glenn Loury for writing
The Anatomy of Racial Inequality and for your constructive criticisms
that made this book stronger. Thanks to David Spergel for reading
and providing scientific guidance on the manuscript. Thanks,
Indradeep Ghosh, for your frank insights and guidance, Joao
Magueijo, for many discussions, and Jaron Lanier, for co-writing a
chapter and teaching me how to think outside the box. Finally, thank
you to members of the Alexander Theory Group at Brown University.
Thank you Jerome Alexander, Salvador Almagro Moreno, Asohan
Amarasingham, Sarah Bawabe, Willis Bilderback, Brown University,
Will Calhoun, Liam Carpenter-Urquhart, Saint Clair Cemin, Colin
Channer, Dwayne Ray Cormier, Everard Findlay, Batia Freedman-
Shaw, Ashok Gangadean, Melving Gibbs, Heather Goodell, Jeff
Greenwald, Leah Jenks, Ned Kahn, Brian Keating, Dagny Kimberly,
Jaron Lanier, Janna Levin, Evan McDonough, Fernanod Pezzino,
Vernon Reid, David Rothenberg, Susan Sharin, Jim Simons, Lee
Smolin, Richard Snyder, Greg Tate, Greg Thomas, and Eric
Weinstein.
Discover Your Next Great Read

Get sneak peeks, book recommendations, and news about your


favorite authors.

Tap here to learn more.


© Heather Goodell

Stephon Alexander is a professor of physics at Brown University


and the 2020 president of the National Society of Black Physicists.
He is also a jazz musician and released his first electronic jazz
album Here Comes Now with Erin Rioux and God Particle with
bassist Melvin Gibbs. The author of Jazz of Physics, Alexander lives
in Providence, Rhode Island.
NOTES
Chapter 2: The Changeless Change
1. Mathematically we can define a space-time vector as
where c is the speed of light and has dimensions
. This gives the time axis a dimension of length since the

dimensions of the time interval, , is length.


2. A local, freely falling observer in an exactly homogenous
gravitational field experiences no forces. However, in a realistic
situation there are gravitational field lines that are inhomogenous
around the observer that will yield tidal forces.
3. Our direct experience of being in rotating environments seems
different from being at rest, especially since rotation is a form of
acceleration, which seems to mimic the effect of a force. We will
soon see that this was part of Albert Einstein’s insight about the
equivalence of acceleration and being subject to a gravitational field
without accelerating.
4. A geodesic can be understood more geometrically with vectors
that are tangent to a point on a curved space, called a tangent
vector. If we take a tangent vector, , and transport it to a nearby
region while keeping it parallel to itself, this amounts to satisfying the
geodesic equation:

5. A solid example of a dynamical equation is that for how a


pathogen like a virus grows in time:
Rate of growth of virus at a later time = (R-1)amount of virus
at earlier time
Or symbolically

= (R-1)y(t)
Here R is famously known as the reproduction number. As I write
this the COVID-19 virus has an average R of 2.3 in the United
States. This leads to a solution of the number of infections to be
exponentially growing in time .
6. All four Maxwell equations are contained in one equation if we
write it in a manifestly Lorentz covariant form: . The left-
hand side of the equation is the four-dimensional derivative of the
field strength tensor that contains all electric and magnetic field
information, and the right-hand side is the four-dimensional current,
containing the electric charge and currents.
Chapter 3: Superposition
1. The concept of phase space was pioneered by Ludwig
Boltzmann, Henri Poincaré, and Josiah Willard Gibbs. Phase space
is a key tool in formulating thermodynamics and statistical
mechanics and in identifying attractors in chaos theory.
2. The particles, like the photon, that communicate the forces
between matter are called bosons and have integer spin. But half-
integer spin belongs in Alice’s Wonderland. To get a feel, consider a
child riding a merry-go-round. If the merry-go-round had half-integer
spin, a child would have to go around two full rotations to ap-pear at
his or her original position. Likewise, an electron needs to spin not
360 but 720 degrees to come back to its original spin orientation. M.
C. Escher’s drawing of ants traversing a Möbius strip is an example
of a space that requires two orbits to return to the same place.
Another fundamental characteristic of an electron is that it never
stops spinning.
3. I even used the concept of quantum entanglement in my role of
science adviser for Disney’s A Wrinkle in Time, directed by Ava
DuVernay.
4. Wave functions are vectors that live in a complex vector space
called a Hilbert space. Said another way, a Hilbert space is a space
of all possible wave functions and has all the usual properties of
linear vector spaces commonplace to linear algebra.
Chapter 4: The Zen of Quantum Fields
1. Kensho is a Japanese word that translates to what Westerners
call enlightenment, a liberating experience recorded by Siddhartha
Gautama, the historical Buddha. According to the Buddha this
potential to experience kensho is accessible to all humans. Similar
accounts of the enlightenment experience were recorded by
individuals across different cultures. Meister Eckhart calls it a
breakthrough. The theologian St. Symeon calls it waking up. “I
wasn’t there.” “There was only the tree.” “A sense of utter liberation
and bliss.” “It is overwhelmingly positive.” “It’s like being drunk, but
on reality.” “It’s more real than real.”
2. This cosmic ocean is called the vacuum state of quantum field
theory. The vacuum state can both create and annihilate particles
with quantum operators called creation and annihilation operators,
respectively.
3. This “simple way” is a linear interaction between the photon
vector potential and the fermion current. Such a linear interaction
guarantees a local phase invariance for the electrons provided the
gauge field simultaneously compensates the phase with its own
phase transformation.
Chapter 5: Emergence
1. In modern times, solid-state physics is often called condensed-
matter physics, which also involves the study of the various states of
matter, such as liquid, gaseous, and crystalline, where many
quantum particles interact quantum mechanically.
2. Kenneth Chang, “When Superconductivity Became Clear (to
Some),” New York Times, January 8, 2008,
https://www.nytimes.com/2008/01/08/science/08super.html.
Chapter 6: If Basquiat Were a Physicist
1. The electron can only have discrete (quantum) spin about its
axis of rotation and never stops spinning. This is to be contrasted
with a macroscopic spinning object, which can continuously change
its spin orientation from up to down, such as when a top starts
spinning and eventually falls to its surface. If the electron’s spin flips
from up to down, it does so discretely, in a quantum unit of spin. And
strangely, experiments seem to require one electron to coexist in a
spin up and spin down state at the same time.
2. A conformal invariant theory is analogous to looking at features
of a theory under a “magnifying glass.” This is equivalent to zooming
in or rescaling the coordinates of the observables of the theory. If the
predictions of the theory do not change under this zooming in, or
zooming out, the theory is said to be scale-invariant. The microcosm
has the same properties as the macrocosm. In some special cases,
conformal invariance is the same as scale-invariance.
3. Thomas Kuhn, The Structure of Scientific Revolutions
(Chicago: University of Chicago Press, 2012).
4. Du Bois wrote that “the Negro is a sort of seventh son, born
with a veil, and gifted with second-sight in this American world,—a
world which yields him no true self-consciousness, but only lets him
see himself through the revelation of the other world. It is a peculiar
sensation, this double-consciousness, this sense of always looking
at one’s self through the eyes of others, of measuring one’s soul by
the tape of a world that looks on in amused contempt and pity. One
ever feels his twoness,—an American, a Negro; two souls, two
thoughts, two unreconciled strivings; two warring ideals in one dark
body, whose dogged strength alone keeps it from being torn
asunder.” W. E. B. du Bois, The Souls of Black Folk (Orinda, CA:
SeaWolf Press, 2020).
5. The conscience collective constitutes the shared social values
within any social order; it regulates norms and activities in every
social order, but it is not always well integrated; the values are not
always consistent, and thus the social order might be poorly ordered.
Here we might think about the conscience collective for both the
larger society and for the community of physicists.
6. Phenomenologists sometimes call these cultural norms, which
differentiate between sense and nonsense, the “lifeworld.” If the
lifeworld is functioning effectively, it is tacit—we are unaware that we
are seeing the world through it.
7. Durkheim recognized three functions of punishment. The first is
the one usually discussed in the literature. Punishment creates an
incentive to conform to social and cultural norms; it realigns what is
in the interest of actors. Second, as noted in the text, punishment
delineates the boundary of what is allowed. If a public violation of a
normative expectation goes unpunished, it will undermine the clarity
of the normative orientation; in contrast, the punishment of specific
actions reinforces normative boundaries. Third is what we call the
“sucker effect.” If conformity to a normative expectation continually
disadvantages an actor, this will undermine her commitment to the
relevant norm; she will feel like a sucker conforming while those who
violate the norm are advantaged over her. If she knows, instead, that
violators are likely to be punished for their violations, this enables her
to sustain her sense of obligation.
8. There is a huge amount of sociological literature on deviance
and marginality. Robert Ezra Park coined the term “marginal man,”
but his usage was different than ours. Also relevant is Georg
Simmel’s discussion of “the stranger.” Everett Hagen was one of the
first to ask if deviance might be related to innovation. See his On the
Theory of Social Change: How Economic Growth Begins
(Homewood, IL: Dorsey Press, 1962), 573.
9. Eric Felisbret, “Legal Venues Celebrate Graffiti as an Art
Form,” New York Times, July 18, 2014,
https://www.nytimes.com/roomfordebate/2014/07/11/when-does-
graffiti-become-art/legal-venues-celebrate-graffiti-as-an-art-form.
Chapter 7: What Banged?
1. The expanding universe model was independently discovered
by
1. Alexander Friedmann et al., “Über die Krümmung des
Raumes,” Zeitschrift für Physik A 10, no. 1 (1922): 377–86.
2. Georges Lemaître, “Expansion of the Universe, A
Homogeneous Universe of Constant Mass and Increasing
Radius Accounting for the Radial Velocity of Extra-Galactic
Nebulæ,” Monthly Notices of the Royal Astronomical
Society 91, no. 5 (March 1931): 483–90.
3. H. P. Robertson, “Kinematics and World Structure,”
Astrophysical Journal 82 (November 1935): 284–301.
4. A. G. Walker, “On Milne’s Theory of World-Structure,”
Proceedings of the London Mathematical Society, Series 2
42, no. 1 (1937): 90–127.
Chapter 8: A Dark Conductor of Quantum Galaxies
1. All observed galaxies have dark matter. See Jim Shelton, “New
Studies Confirm Existence of Galaxies with Almost No Dark Matter,”
YaleNews.com, March 29, 2019,
https://news.yale.edu/2019/03/29/new-studies-confirm-existence-
galaxies-almost-no-dark-matter.
2. The spatial anisotropy in the density of baryonic and dark
matter deviates by approximately one part in ten thousand from the
average homogeneous and isotropic cosmic microwave background
thermal bath.
3. It was shown by Kris Pardo and David Spergel that it is difficult
to reproduce the effects of dark matter in the CMB fluctuations in the
early universe with MOND. See Kris Pardo and David Spergel,
“What Is the Price of Abandoning Dark Matter? Cosmological
Constraints on Alternative Gravity Theories,” Physical Review
Letters 125 (October 19, 2020): https://arxiv.org/abs/2007.00555.
4. To be precise, dark matter is five times the density of baryonic
(visible) matter.
5. Kaplan is known for producing the critically acclaimed film
about the Large Hadron Collider, Particle Fever, which tells the story
of the vision of the LHC and the physicists behind it. Coincidentally, I
was the moderator for the world premier for the film at the
NeueHouse in New York City.
6. Classical field theories have conserved quantities. For
example, in electromagnetism the current is conserved. A quantized
theory should also conserve all classical currents. However, there
are some currents that are broken upon quantization. The amount of
current that is broken is called an anomaly. However, a healthy
quantum theory fixes these anomalies by realizing that topology
needs to be considered in the quantum theory so as to reinstate the
current conservation.
Chapter 10: Embracing Instabilities
1. The uncertainty principle can be conveniently written as
, the uncertainty in the position of the particle in a harmonic
oscillator is bounded by the size of the system. Therefore, the
velocity can never be zero according to the uncertainty relation
above.
2. General relativity provides an equation that relates energy and
matter to the curvature of space-time. The configuration of the matter
and energy should warp the space-time field in a specific dynamical
manner dictated by the Einstein field equations similar to how a
magnet emanates and bends a magnetic field surrounding it.
Microscopically, a magnet emerges from interacting atomic spins in a
metal. The cosmological constant interacts with the space-time in a
manner that causes it to accelerate in its expansion without bound.
3. The Ashtekar variable is similar to QCD in that the dynamics of
both theories are encoded in a gauge field connection. In QCD the
connection enjoys a larger symmetry [SU(3)] than that Ashtekar
connection, [SU(2)].
4. Stephon Alexander and Raúl-Rubio, “Topological Features of
the Quantum Vacuum,” Physical Review D 101, no. 2 (2020);
Stephon Alexander, Gabriel Herczeg, and Jinglong Liu, “Chiral
Symmetry and the Cosmological Constant,” Physical Review D 102,
no. 8 (2020).
Chapter 11: A Cosmologist’s View of a Quantum
Elephant
1. I first spoke to Sir Roger Penrose, the inventor of twistor theory,
who encouraged me to discuss the idea with Ashtekar. Twistors are
maps from events in space-time to the celestial sphere that has
close semblance to the symmetries enjoyed by the Ashtekar
connection. I first thought that my idea behind parity violation and
quantum gravity had to do with twistors.
Chapter 12: The Cosmic Biosphere
1. “John von Neumann Compares the Functions of Genes to Self-
Reproducing Automata,” HistoryofInformation.com,
http://www.historyofinformation.com/detail.php?id=682.
2. B. Jesse Shapiro et al., “Origins of Pandemic Vibrio cholerae
from Environmental Gene Pools,” Nature Microbiology 2, article no.
16240 (2017).
As Salvador Almagro-Moreno communicated to me, “Virulence
adaptive polymorphisms (VAPs) circulate in environmental
populations and must be present in the genomic background of a
bacterium before it can emerge as a successful pandemic clone.”
3. The entropy of a black hole is calculated to be .
4. Ludwig Boltzmann, “The Second Law of Thermodynamics,” in
Theoretical Physics and Philosophical Problems, ed. B. F.
McGuiness (New York: D. Reidel, 1974).
5. One potential loophole to the low-entropy initial state is that
there was a previous state what was actually even more entropic
that reduced its entropy. For example, if the big bang started out
dominated with black holes, then they could evaporate and produce
the CMB radiation. If the massless radiation, such as photons, got
homogeneously distributed then decayed into matter, then we would
have a lower-entropy situation. However, the original entropy
presumably in the form of black holes would have to get diluted. This
idea can perhaps be implemented in a mechanism developed by
Peter Mészáros and me where we postulated that primordial black
holes that form after inflation could explain the CMB and dark matter
if they undergo Hawking evaporation.
6. Fred C. Adams et al., “Constraints on Vacuum Energy from
Structure Formation and Nucleosynthesis,” JCAP 03 (2017).
Chapter 13: Dark Ideas on Alien Life
1. A similar phenomenon happens with electrons and magnetic
fields in the fractional quantum Hall effect.
Chapter 14: Into the Cosmic Matrix
1. Arvind Borde and Alexander Vilenkin demonstrated that even
cosmic inflation is geodesically incomplete and does not escape the
cosmic singularity. See Arvind Borde, Alan H. Guth, and Alexander
Vilenkin, “Inflationary Spacetimes Are Incomplete in Past Directions,”
Physical Review Letters 90, no. 15 (April 15, 2003).
2. To be precise, if x and p are position and momenta variables
then quantization rules promote their Poisson brackets (curly
brackets) to commutation (square brackets) . Classical
dynamics of an observable is given by Hamilton’s equation, which
states that the time evolution is generated by the Poisson brackets of
the Hamiltonian and the observable of interest. For example, the
classical time evolution of the position of a particle, x(t) is given by
the Poisson bracket between the Hamiltonian and the position.

Whereas the time evolution of the position operator is given by


the commutator between the Hamiltonian and position operators.

3. Here we are assuming that the spatial dimensions have the


topology of circles whose local products are tori. These topologies
admit one-cycles that strings can wind around without collapsing to a
point.
4. This initial condition for string cosmology is an extension of the
Copernican principle in the standard big bang, which states that
there is no preferred vantage point in the universe. The Copernican
principle is consistent with assuming homogeneity and isotropy not
only in space but in the degrees of freedom that occupy the
universe.
5. A cyclic BV mechanism was studied by Brian Greene, Daniel
Kabat, and Stefanos Marnerides. They found that the BV mechanism
can exhibit a cyclic cosmology if the theory incorporates higher
derivative terms in the gravitational theory.
6. In a closed string theory we can add open strings that satisfy
both von Neumann and Dirichlet boundary conditions.

The above equation states that the derivate of the string ends on
p space-time dimensions. This is a Dirichlet boundary condition. This
p-dimensional hypersurface defines the worldvolume of a Dp-brane.

7. The above equation states that the string vanishes in the other
dimensions.
Chapter 15: The Cosmic Mind and Quantum
Cosmology
1. Deepak Chopra, MD, and Menas C. Kafatos, PhD, You Are the
Universe: Discovering Your Cosmic Self and Why It Matters (New
York: Harmony Books, 2017).
2. Kimbwandende Kia Bunseki Fu-Kiau, African Cosmology of the
Bantu-Kongo: Principles of Life and Living (Brooklyn, NY: Athelia
Henrieta Press, 2001), 17–54.
3. In general relativity the Hamiltonian is constrained to be zero.
As a result the associated Schrödinger equation will no longer have
time dependence. The state of the universe will therefore be
timeless, which means that the universe as a whole is changeless. In
order to obtain time dependence, one has to introduce an external
clock relative to the rest of the universe, and matters of this sort are
called the problem of time in quantum gravity.
4. Alexander Vilenkin, “Creation of Universes from Nothing,”
Physics Letters B 117, no. 1–2 (November 1982): 25–28.
5. Andrei Linde, “Universe, Life, Consciousness,”
https://static1.squarespace.com/static/54d103efe4b0f90e6ca101cd/t/
54f9cb08e4b0a50e0977f4d8/1425656584247/universe-life-
consciousness.pdf.
6. Vincent Jacques et al., “Experimental Realization of Wheeler’s
Delayed-Choice Gedanken Experiment,” Science 315, no. 5814
(February 16, 2007): 966–968, https://arxiv.org/abs/quant-
ph/0610241.
ALSO BY STEPHON ALEXANDER
The Jazz of Physics
PRAISE FOR FEAR OF A BLACK UNIVERSE

“The rabbit hole gets wrestled here. An old school saying applies:
the more you know, the more you don’t know. Dance along this
read into the unknown and find out that this book may be the best
ever answer to ‘what is soul?’”
—Chuck D, rapper and co-founder of Public Enemy

“This book reminds me of Hawking’s A Brief History of Time—


very brief and very ambitious. It covers an enormous amount of
material and offers insights not only into physics but how we do
physics and who we are as physicists.”
—David Spergel, winner of the 2018 Breakthrough Prize in
Fundamental Physics

“Stephon Alexander has done it again. Fear of a Black Universe


opens many dimensions—it’s an endlessly stimulating, hyper-
complex overview by a deeply musical scientist and
mathematician. From Public Enemy’s classic Fear of a Black
Planet hip hop album and what happened before the Big Bang to
how consciousness itself is woven into the fabric of space-time,
this book will blow your mind. A must-read for anyone who thinks
of physics and music as being inseparable.”
—Paul D. Miller aka DJ Spooky

“An expansive and poetic account of not just the theory of


physics, but the dreamy processes that lead to its creation, and
the opposing forces that support and hinder its progress.”
—Eugenia Cheng, author of x + y

“Einstein famously remarked that mystery is the source of all true


art and science. This book explores some of the biggest
mysteries of all—dark matter, dark energy, the origin of the
universe, and the origin of life—in ways that are unconventional
and enthralling, yet down to earth. We go on a journey with a
brave adventurer for whom physics is a passionate pursuit of
beauty and truth. His passion shines through on every page.”
—Edward Frenkel, author of Love and Math

“Read this book and you’ll feel awe at the grandeur and the
remaining mysteries of our world, but you’ll also get a hint of the
human side of physics. Science is made of people and is for
people; this book revives the humanist project that launched
science in the first place.”
—Jaron Lanier, author of Ten Arguments for Deleting Your Social
Media Accounts Right Now

“In this courageous and provocative book, Alexander recounts his


personal story of overcoming prejudice while offering a hopeful
perspective for our future. Discussing the origins of his boldest
ideas, from his practice as a professional jazz musician to his
explorations of Jungian psychology, is especially inspiring.”
—Lee Smolin, author of Einstein’s Unfinished Revolution

You might also like