Fear of A Black Universe - Stephon Alexander
Fear of A Black Universe - Stephon Alexander
Fear of A Black Universe - Stephon Alexander
Hachette Book Group supports the right to free expression and the
value of copyright. The purpose of copyright is to encourage writers
and artists to produce the creative works that enrich our culture.
Basic Books
Hachette Book Group
1290 Avenue of the Americas, New York, NY 10104
www.basicbooks.com
The publisher is not responsible for websites (or their content) that
are not owned by the publisher.
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.
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.
SUPERPOSITION
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.
EMERGENCE
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.
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:
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
WHAT BANGED?
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.
A DARK CONDUCTOR OF
QUANTUM 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
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.
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.
A COSMOLOGIST’S VIEW OF A
QUANTUM ELEPHANT
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.
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.
= (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.
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
“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