MY EINSTEIN [8.15.06]
Essays by Twenty-Four of the World’s Leading Thinkers on the Man, His Work, and His Legacy

Edited by John Brockman

"RELATIVELY FASCINATING" — The Washington Post



"IRRESISTABLE" — The Buffalo News

"Relatively Fascinating"

"From the vaudevillian ('Einstein, Moe, and Joe') to the tantalizing ('The Greatest Discovery Einstein Didn't Make')...My Einstein delivers even more than its lengthy title promises." — Washiington Post Book Review

"Irresistible"

These essays are irresistible ... the charm of the book is that its often star-struck writers so freely wanted to be connected to entirely non-theoretical humanity, their own and Einstein's.


(Salt Lake City)

"A Gem"

My Einstein is a gem of a book that celebrates not only Einstein the scientist but also Einstein the man, even though it is a collection of essays written by scientific figures ... The result is a remarkably well-rounded figure.

"Excellent"

This excellent collection features 24 of the world's leading theoretical and experimental physicists, cosmologists, science historians and science writers as they examine this incredible man and his work.




Table of Contents

Introduction [Excerpted]
The fights we got into were almost always part of a broader history lesson: Philip and I discovered that we were personally responsible for the death of the Second Person of the Holy Trinity. We tried reasoning. None of our arguments — Jesus was a rabbi, who prayed in Hebrew and preached in a synagogue; his mother looked like our mother, not like their mothers — seemed to impress these furious young Irishmen. — John Brockman
Editor & Publisher, Edge; Author, By the Late John Brockman; Editor, Intelligent Thought: Science Versus the Intelligent Design Movement

Einstein When He's at Home
The popular image of Einstein as archetypal eccentric boffin dates to half a century after the first flowering of his astonishing creative genius. The tangle-haired sage who has launched a thousand brands, whether of computer or coffee mug or T-shirt, is an Einstein who is well past his scientific best, a faded version of the original. We should bury the sockless dustball who rolled around Princeton and restore the creative Einstein. — Roger Highfield

Science Writer; Science Editor, The Daily Telegraph
; Author, The Science of Harry Potter: How Magic Really Works

The Freest Man
By the 1920s, with the experimental confirmation of the general theory of relativity by the observed bending of light and the Nobel Prize for his work on quantum theory, Einstein was probably the world's preeminent scientist, certainly its leading theoretical physicist and the inspiration for a rising generation. However, from 1927 on, he increasingly distanced himself from his fellow physicists by his unwillingness to accept what became their central belief—the Copenhagen interpretation of quantum mechanics. There was a great deal of sadness for them in this—particularly because Einstein had contributed as much or more than any individual in bringing about the revolution spearheaded independently by Werner Heisenberg and Erwin Schrödinger….Yet my sense is that Einstein never minded this scientific isolation. Like the other ties he was willing to give up, the consensus in physics could also be dispensed with….This apartness is reflected in his work, lending it a strange quality of permanence, a weightiness that is absent in efforts by others.
Gino C. Segrè
Theoretical physicist, University of Pennsylvania; Author, A Matter of Degrees: What Temperature Reveals About the Past and Future of Our Species, Planet, and Universe

Mentor and Sounding Board
My geon paper was mostly classical (i.e., non-quantum) but it contained a few remarks about quantum physics, enough to elicit comment from Einstein. He told me once again, as he had so often in the past, that he did not like the probabilistic nature of quantum theory. Nearly fifty years had passed since he introduced that prototypical quantum entity, the photon (as we now call it). He couldn't stop thinking about, and worrying about, the quantum world that he helped bring into being. Now, in my own later years, I find myself pondering quantum theory, too, with one of my favorite questions, "Why the quantum?" There is something about quantum theory that is rightly more troubling than relativity, something still calling out for a deeper explanation.
John Archibald Wheeler
Physicist, professor emeritius, Princeton and University ot Texas; Coauthor (with Kenneth W. Ford) Geons, Black Holes, and Quantum Foam: A Life in Physics

My Einstein Suspenders
Aesthetic arguments, while useful as development tools, especially when there are no observations to guide the effort, made me uneasy—seemed a throwback to Greek reasoning about the celestial spheres. More recently, I came to realize that Einstein based special relativity not on pure thought alone but upon a great deal of physical observation and codifying theory—in particular, electromagnetism and the theory of light via James Clerk Maxwell's equations. Einstein was certainly aware of Lorentz's work, but was coming from the Maxwell side, not the Michelson-Morley results. He was reducing these ideas down to two essential postulates added onto the existing physics: (1) The speed of light is definite and independent of the speed of the source or of the observer, and (2) the laws of physics are the same in every inertial frame. From these two postulates and thought experiments, one can derive all the consequences of special relativity, including the Lorentz transformations, time dilation, length contraction, loss of simultaneity, E=mc2, and the lot!
George F. Smoot
Physicist, University of California, Berkeley; Author, Wrinkles in Time

Einstein, Moe, and Joe [Excepted]
Einstein was examining patents during the long days and presumably working on physics nights and weekends—why? He had not been driven by some experimental breakthrough (although there were growing experimentally inspired doubts about the Newtonian worldview) but by an aesthetic and deep physical sense of the accordance of symmetry with nature....Who could not love the iconoclast who blew up something called "the luminiferous aether"?
Leon M. Lederman
Physicist; Director Emeritus, Fermi National Accelerator Laboratory; Nobel Laureate; Author, The God Particle

The True and the Absurd
Einstein could not simply accept a theory that was handed to him. It was not in his nature, for he was the jujitsu master of physics. Armed only with intellect and the weapon of the gedankenexperiment—the thought experiment—Einstein had an unparalleled ability to overthrow a theory by using its own power against it. The stronger the theory, the more subtle and dangerous the gedankenexperiments; with his reductio ad absurdums, he laid naked the contradictions in the commonsense picture of the universe.
Charles Seife
Science Writer and Professor of Journalism, NYU; Author, Zero: The Biography of a Dangerous Idea

Albert Einstein: a Scientific Reactionary
Einstein later remarked that once he realized that Newton's idea of absolute time was suspect, he was able to work out within six weeks how to modify Newton's mechanics to make it consistent with Maxwell's equations. Because there are revolutionary implications to Einstein's mechanics—E = mc2 being the most familiar—it is not often realized how profoundly conservative Einstein's innovation actually was. It was a minimal modification of the fundamental physics equations of his day. Maxwell's equations have a fundamental speed—the speed of light—built into them in an essential way. Removing this speed would require a major reworking of the equations. By comparison, it was a trivial matter to change Newton's mechanics to include this speed limit, using light signals to coordinate the time measurements of separated clocks, an idea Einstein picked up from his day job as a Swiss patent examiner. Changing Maxwell's equations to make them consistent with Newtonian mechanics would almost certainly have ruined their agreement with experiment, whereas the changes Einstein made in Newton's mechanics would show up only at speeds comparable with that of light. Frank J. Tipler
Mathematical physicist, Tulane ; Author,The Physics of Immortality

Helen Dukas: Einstein's Compass
I was seven years old, and my sister Esther eight, when Helen Dukas, who had been Einstein's personal secretary since 1928 and his literary executor since his death in 1955, began making regular weekly visits to baby-sit for us and a growing brood of younger sisters at the Dyson household on Battle Road in Princeton…."Helen could remember infallibly who had written what when, who needed an answer and who didn't, who was an earnest seeker after truth, and who was a journalistic pest," my father recalled at her memorial, adding that her presence allowed Einstein "to live the life of an absent-minded professor; she kept to herself the tiresome details that he wanted to forget, and she reminded him of the important things he wanted to remember." To the rest of the world, Einstein achieved immortality through his science, his humanity, and the celebrity he enjoyed while alive. To friends and neighbors in Princeton, Einstein achieved immortality through Helen Dukas. My sisters and I were too young to have known Einstein, but Helen's weekly visits brought him back to life for us.
George Dyson
Science Historian; Designer; Author, Project Orion

My Three Einsteins
Einstein's much repeated use of the word "God" was not an indulgence and not a purely symbolic act. It was a well-considered philosophical position. He acknowledged that a truly universal theory of physics has theological implications; at the same time, he worried intensely about the destructive power of religions whose adherents imagine they can pray for their success or for others' failure. Einstein believed, passionately if a bit naïvely, that his logical approach could help here, too. "After religious teachers accomplish the refining process indicated, they will surely recognize with joy that true religion has been ennobled and made more profound by scientific knowledge," he wrote in 1941. — Corey S. Powell
Science Writer; Editor, Discover magazine; Adjunct Professor, New York University; Author, God in the Equation: How Einstein Transformed Religion

In Search of Einstein
I found in the library the report of the Solvay Conference of 1927, with transcripts of the debates between Einstein and Bohr and their discussions with their colleagues on the quantum theory, and I read every word carefully. I found Bohr's reasoning fascinating but in the end unconvincing. Einstein by that time had persuaded me that quantum mechanics is incomplete and requires replacement by a new theory, and this is still my view….Although I respect my colleagues who disagree, I find their thinking basically incomprehensible....Did the universe wait almost 14 billion years for the descendants of the ape to decide to do experiments before its wavefunction collapsed? Is the world just information waiting to be decoded? I have worked with quantum mechanics all my life and it still makes as little sense to me as it did the first year I learned it. So I take some small comfort in the fact that it never made sense to Einstein, either. Lee Smolin
Founding Member, Research Physicist, Perimeter Institute for Theoretical Physics, Waterloo, Ontario; Author: The Trouble With Physics

Einstein and Absolute Reality
The discovery that (trivial exceptions aside) quantum physics makes only probabilistic predictions is certainly one of the deepest philosophical discoveries of science. After all, the program of science over the centuries has been investigatio causarum, the investigation of causes. And after centuries of digging deeper and deeper along the causal chain, we finally came to a stop. The individual quantum event happens by chance. There is no hidden cause, no hidden reason. But fundamental randomness is unbearable to us…. Einstein was disturbed by this. He supposedly once exclaimed that if that randomness remained with us, he would rather work in a casino than as a physicist. Anton Zeilinger
Experimental Physicist, University of Vienna; Author, Einsteins Spuk. Teleportation und weitere Mysterien der Quantenphysik
(forthcoming in English: Quantum Teleportation)

A Walk Down Mercer Street
Over the next few years of high school, I read everything I could about physics and math….I'd curl up in a big soft chair in my high school library… captivated by a compendium of essays called The World of Mathematics, a four-volume set of reprinted articles by geniuses like Poincaré, Newton, and Bertrand Russell…. Meanwhile I kept reading about Einstein. I liked his simplicity and his determination to think for himself, to take on the giants who had preceded him. I especially admired his cockiness. When asked how he would have felt if Arthur Eddington's eclipse observations had not confirmed his prediction, based on general relativity, that starlight would be bent by gravity as it passed by the sun, Einstein is said to have replied, "I would have been sorry for the dear Lord; the theory is correct." At this stage in my life I wasn't able to understand Einstein's scientific ideas in any depth, but perhaps that didn't matter so much. What I cared about more were the other lessons he taught—about how to act as a scientist, how to feel about God and authority and the wonder of the universe, how to fight, how to be stubborn, how to trust your instincts, and how to admit when you're wrong. — Steven Strogatz
Physicist, Cornell University; Author, Sync: The Emerging Science of Spontaneous Order and the best-selling textbook Nonlinear Dynamics and Chaos

Things and Thoughts
During my undergraduate years, for one summer at the Institute for Advanced Study in Princeton, I worked on the Einstein Papers publication project, which was just then getting under way. I found it extraordinary to see how deeply Einstein had been engaged with detailed discussions of inventions and patents. For my PhD thesis in the history of science…I used the case of Einstein's work on the gyrocompass—a nonmagnetic way of tracking one's orientation—to show how technological concerns, the grit of the basement, lay behind some of Einstein's most abstract thought experiments. The gyrocompass became for Einstein a model of the atom. Pure physics met applied engineering. — Peter Galison
Mallinckrodt Professor of History of Science and of Physics, at Harvard University; Author, Einstein's Clocks, Poincaré's Maps

Childe Bernstein to Relativity Came
I did not have any idea what it meant to "understand" a physics theory like relativity. The kind of understanding I was familiar with from high school involved being able to translate a foreign language like Latin into English; having done that, one understood the Latin. Understanding geometry meant being able to repeat the steps of a proof on an exam. Understanding a poem meant understanding, perhaps with the aid of a dictionary, all the words and allusions in it….I assumed that understanding relativity was something like this. I would find a book and, with the aid of a dictionary, translate all the unfamiliar words into ones that I understood. I was prepared, if necessary, to devote a couple of months to this project. — Jeremy Bernstein
Emeritus professor of physics, Stevens Institute of Technology.; Author, Oppenheimer: Portrait of an Enigma

The Books in the Basement
These insights—the insights of an amateur—fade from disuse, only to be rekindled every few years as I open a new book on Einstein and take in another production of the metaphorical stage play. The trains and the lightning bolts, the elevator and the light beam—coming upon them is like encountering old friends. With each retelling, the ideas settle in a little more comfortably….In the relativistic universe, all motion is shared among four dimensions. As I sit at my desk going nowhere, I am moving full speed ahead through time. If I get up and start walking, my spatial velocity must be subtracted from my temporal velocity. My watch runs incrementally slower and I don't age quite so rapidly. George Johnson
Science Writer, New York Times; Author, Miss Leavitt's Stars

How He Thought
It's the mind that fascinates me—a way of thinking that I admire above all others. In that year, whose hundredth anniversary we have just celebrated, Einstein was at the height of his powers. He had an almost supernatural way of looking into nature and seeing clearly what others could see only as cloud-shrouded shadows. Not that he could decipher unusually complicated formulas, digest difficult mathematics, or remember prodigious amounts of experimental information. Einstein's style was to begin with the simplest observations about nature—things so simple even a clever child could understand them. But from these elementary considerations, he drew the most profoundly far-reaching conclusions. The things he saw were in retrospect obvious, but no one else had seen them.
Leonard Susskind
Physicist, Stanford University; Author, The Cosmic Landscape

Toward a Moving Train
The story goes that Einstein liked to sleep ten hours a night—unless he was working very hard on an idea; then it was eleven. And while he slept the night and part of each day away, he dreamed. He dreamed of riding his bike through trees and catching the light as it fell off the leaves. He dreamed of time standing still as he traveled at the speed of light. He dreamed of relativity. He dreamed of curved spacetime.
— Janna Levin
Theoretical physicist, Barnard College; Author, A Madman Dreams of Turing Machines

Einstein's Tie
Einstein's science was a raging iconoclasm, demolishing the very Newtonian notions of absolute space and time that were so cozy and nonthreatening. His science opened the door to an unknown world, one beyond sensory perception—an invisible world with mysterious properties and bizarre effects. Once you stepped into this new worldview, you couldn't go back. Like the mythic hero returning from his quest, you'd emerge transformed, with a new conception of reality. This was science as a rite of passage, science as spiritual fulfillment. Newton's ideas may well have had a similar impact on the minds of early eighteenth-century natural philosophers, because they also revealed invisible connections between the heavens and the earth….Still, Newton's science dealt with palpable reality, while Einstein's went beyond. A different icon for a different age. Marcelo Gleiser
Physicsist, Dartmouth; Author, The Dancing Universe: From Creation Myths to the Big Bang

The Greatest Discovery Einstein Didn't Make
When I look at Einstein's equations, "expansion" sort of screams out at me. Even skeptical students accept an expanding—or contracting—universe as an implication of Einstein's theory of gravity. But for over a decade after Einstein developed his theory, he could not hear what his own equations were saying. I have often wondered how Einstein missed this one. How did Einstein miss the opportunity to predict the expansion of the universe? — Rocky Kolb
Physicist; Director, Particle Astrophysics Center, FermiLab; Author, Blind Watchers of the Sky

The Gift of Time
Time is a remarkably elusive concept. Some treat it as a mere coordinate, a way to help specify an event. If you do so with three spatial coordinates (x, y, z) then time becomes the "fourth dimension"—but only in a trivial sense. In this manner, it appears in most physics equations. But although physics uses time, it is our dirty little secret that we don't really understand time. Physicists will tell you that time is now "unified with space" (thanks largely to Einstein), and we are supposed to be happy with that. But time behaves in a fundamentally different manner from space, in a way that physics doesn't quite acknowledge. Time is significantly more mysterious than space. — Richard A. Muller
Physicist, University of California, Berkeley; Author, Nemesis: The Death Star

Flying Apart
Ironically, the proposal that Einstein himself regarded as his "greatest blunder" might have been right all along. This was a late modification—sometimes unkindly called a fudge factor—that he made to the crowning achievement of his career, the general theory of relativity. I became fascinated with Einstein's fudge factor when I was a student in the 1960s. Unfashionably, I found it tantalizing rather than repugnant, and over the years I have argued in its favor in the face of widespread contempt for it. Now the tables are turning, and scientists are reluctantly admitting that maybe Einstein was wrong to think he was wrong. — Paul C. W. Davies
Phyicist, Australian Centre for Astrobiology, Macquarie University, Sydney; Author, How To Build A Time Machine

Einstein in the Twilight Zone
We currently have no idea what might be responsible for the observed acceleration of the universe, but the best bet is something like a cosmological constant. We now understand this term in a different way than Einstein did. It turns out that if one allows empty space to have energy, then this will automatically result in the appearance of a cosmological constant. And the laws of quantum mechanics, when combined with relativity, imply that such a term should be present—that is, that empty space should have energy. The only problem is that when we try to estimate how much energy empty space should have, we derive a number about 120 orders of magnitude larger than is allowed by observations. Clearly, there is something profound that we do not yet understand at the interface of quantum mechanics and gravity. Whether or not it will require a fundamental revision of our understanding of the theory that Einstein discovered, it is likely that the ideas he introduced will be at the heart of the matter. — Lawrence M. Krauss
Ambrose Swasey Professor of Physics, Case Western Reserve University; Author, Hiding in the Mirror

No Beginning and No End
What irony that dark energy has been revived and Einstein's intuition on this count has been vindicated, but in a context so antithetical to his original dream! Or could it be that the discovery of dark energy has deeper significance? Could this be a sign that Einstein was closer to the truth in 1917 than we are today? I find myself asking these questions because during the last few years Neil Turok, at Cambridge University, and I have been developing a new competitor for the hot Big Bang model….When we started down this path, we did not know where we were headed, and the chances of success seemed minuscule, given the wealth of new astronomical observations that had ruled out all previous competitors. Nevertheless, we persisted and found a surprisingly simple, logical alternative we call the cyclic model. As it turns out, without intending to, we found an alternative to the hot Big Bang picture that is just as effective in explaining the universe we observe and yet comes closer to embodying Einstein's vision. Paul J. Steinhardt
Albert Einstein Professor in Science, Princeton University; Author, The Cyclic Universe (forthcoming)

Where Is Einstein?
And where is Einstein today?... In those who take sides. In every anti-nuclear-weapons demonstration. In all young people who exhibit a slight disrespect for authority and all old people in authority who are radicals. In those who have fled from tyrannical and oppressive families and narrow-minded educators. In all our students who can solve a problem we cannot. In all the people we think of as innocent geniuses who take themselves not very seriously. In all stubborn and difficult physicists who think of physics problems as a reason for being. And in all those who tackle problems they know they can't or won't solve. Maria Spiropulu
Experimental physicist, CERN


Dedication

To Sidney Coleman, a true seeker


Introduction
John Brockman

Readers of this book will already know quite a bit about Albert Einstein, whose centennial we celebrated in 2005—the year not of his birth but of his "annus mirabilis," when he produced five papers that have forever altered our perception of reality.

But to reprise the basic facts: Einstein was born on March 14th, 1879, in Ulm, Württemberg, Germany, and died on April 18th, 1955, in Princeton, New Jersey.  The five 1905 papers are his University of Zurich doctoral dissertation on the determination of molecular dimensions and the four more famous ones, listed here in order of their submission to Annalen der Physik:

● on light quanta and the photoelectric effect ("On a Heuristic Point of View About the Creation and Conversion of Light"—this is the work for which he was awarded the Nobel Prize in 1921);

● on Brownian motion ("On the Movement of Small Particles Suspended in a Stationary Liquid Demanded by the Molecular-Kinetic Theory of Heat"); 

● and two papers on special relativity ("On the Electrodynamics of Moving Bodies" and "Does the Inertia of a Body Depend on its Energy Content?" in which appears his famous equation of matter and energy, E = mc2).

In the years following this spectacular production, Einstein devoted himself chiefly to incorporating the gravitational force into his theory of relativity, and in 1916 published "The Foundations of General Relativity Theory." he cosmological constant, later repudiated by him as his "greatest blunder" but now very much back in favor with some cosmologists as a means of describing the recently discovered acceleration of the universal expansion. Einstein was clearly the most important person of the twentieth century. He achieved an iconic status that (some would say unfortunately) transcends even the heights of his scientific genius. 

I have therefore asked the contributors to My Einstein to address the following questions:

The two dozen essayists in My Einstein are among the world's leading theoretical and experimental physicists, science historians, and science writers. But this is not a just a book about physics.  It is a collection of personal narratives, providing a unique window into how these thinkers assess Einstein's scientific and philosophical legacy and his particular influence on their own lives and work. They are:  Who was Einstein to you? What difference did he make to your worldview, your ideas, your science? How did Einstein influence you personally? Who is your Einstein?

The two dozen essayists in My Einstein are among the world’s leading theoretical and experimental physicists, science historians, and science writers. But this is not a just a book about physics.  It is a collection of personal narratives, providing a unique window into how these thinkers assess Einstein’s scientific and philosophical legacy and his particular influence on their own lives and work. They are:

Roger Highfield on the Einstein myth;

John Archibald Wheeler (the only one who actually knew Einstein, though the Nobel laureate Leon Lederman once met him briefly) on their meetings in Princeton, Wheeler on the Princeton physics faculty and Einstein at the Institute for Advanced Study; 

Gino C. Segre, Lee Smolin, and Anton Zeilinger on Einstein's difficulties with quantum theory;

George F. Smoot and Peter Galison on Einstein's blending of pure thought and physical observation;

Leon Lederman on the special theory of relativity;

Charles Seife on Einstein's use of gedankenexperiment;

Frank J. Tipler on why Einstein should be seen as a scientific reactionary rather than a scientific revolutionary;

George Dyson on growing up in Princeton and his friendship with Helen Dukas, Einstein's longtime amanuensis;

Corey Powell on the philosophical underpinnings of Einstein's use of the word "God";

Steven Strogatz, George Johnson, and Jeremy Bernstein on how Einstein turned them on to physics in their early years;

Leonard Susskind on Einstein's way of thinking;

Janna Levin and Maria Spiropulu on how he is perceived among physicists in academe today;

Marcelo Gleiser on Einstein's new world of mysterious properties and bizarre effects;

Paul C.W. Davies, Lawrence Krauss, and Rocky Kolb on the acceleratedexpansion of the universe and the revival of Einstein's cosmological constant;

Richard A. Muller on the mysterious nature of time;

Paul J. Steinhardt on a new cosmology involving a cyclic universe and its relation to Einstein's cosmological thought.


And me? Who is my Einstein?

I remember the moment I found out about Einstein's death, brought up short by a headline at a kiosk in an underground station of Boston's MTA. I was fourteen at the time. It was a shattering moment, in which I felt genuine grief and loss.

By then my family had moved  to the relative peace and quiet of the suburbs, but the first ten years of my life had been marked by learning survival tactics in the "other" Boston—miles away from the graceful sailboats on the Charles River, the gleaming golden dome of the State House on Beacon Hill, the serene beauty of Harvard, the bold architecture of MIT.

I grew up in Dorchester in the 1940s. It was a tough, gritty neighborhood, where, before World War II, Father Charles E. Coughlin, the infamous "Radio Priest," had regularly sent sound trucks up and down the streets spreading his anti-Semitic gospel. This assault had helped to turn Dorchester into a battleground between the Irish kids and the greatly outnumbered Jewish kids. Our three-block walk to the William E. Endicott School on Blue Hill Avenue was a daily obstacle course — my brother Philip, three years my senior, having to defend himself while also protecting me. Our sense of perilous vulnerability was heightened by the realization that anyone with any kind of civic authority—be it a teacher, trolley-car conductor, or cop—seemed always to have a name like Flaherty, O'Reilly, or McCormack. 

The fights we got into were almost always part of a broader history lesson: Philip and I discovered that we were personally responsible for the death of the Second Person of the Holy Trinity. We tried reasoning. None of our arguments — Jesus was a rabbi, who prayed in Hebrew and preached in a synagogue; his mother looked like our mother, not like their mothers — seemed to impress these furious young Irishmen.

But we did have a secret weapon—the most powerful kind, one we realized they would never possess, or even understand. On more than one occasion when we limped home from battle, while tending to our bloody noses, cuts, and scrapes our mother would buck us up, vigorously fighting bigotry in kind;

"Look at them! What the hell do they have? They bake a ham on Sunday and eat it all week!  The men don't bathe!  The women leave their babies in carriages outside the bars! But look what we have!" Her blue eyes beamed strength, certainty, and pride as she dabbed at our bruises. "What we have, they will never have. We have…Einstein!"

My mother was right.  We had Einstein with us, as we made our way up through the terrifying school system and investigated what the public library had to offer. He gave us permission to think big thoughts, to explore intellectually the remotest corners of existence. He allowed us to appreciate, to embrace, the life of the mind. He was always with us. We did have Einstein; we still have Einstein.

My brother Philip become a research physicist and recently retired after a long  career at NASA.  He is now Distinguished Research Associate at NASA and  a recipient of its Exceptional Service Medal. As for me, today I am fortunate to work with, and count among my friends, leading cosmologists, particle physicists, and string theorists, all of them to some degree Albert Einstein's heirs. You could say that I'm very lucky…but maybe luck had nothing to do with it. You see, I had Einstein—my Einstein.

JB


Einstein was examining patents during the long days and presumably working on physics nights and weekends—why? He had not been driven by some experimental breakthrough (although there were growing experimentally inspired doubts about the Newtonian worldview) but by an aesthetic and deep physical sense of the accordance of symmetry with nature....Who could not love the iconoclast who blew up something called “the luminiferous aether”?  Could any of the great scientists of the new century—Poincaré, Lorentz—could any of them have created this idea?  Have some fun: Raise that issue at the faculty club near the physicists’ table and then avoid the flying debris and lurid language as the physicists’ table erupts.

EINSTEIN, MOE, AND JOE
By Leon M. Lederman

LEON M. LEDERMAN, director emeritus of the Fermi National Accelerator Laboratory, received the Nobel Prize in physics in 1988 (with Melvin Schwartz and Jack Steinberger). He is the author of several books, including (with Dick Teresi) The God Particle: If the Universe Is the Answer, What Is the Question?

Leon M. Lederman's Edge Bio Page


EINSTEIN, MOE, AND JOE

It is difficult to convey—even to the most scientifically oriented of lay readers—the awe one scientist feels for another who has done something truly spectacular. If we examine the Gaussian spectrum of physicists, extending from just-barely-made-PhD all the way to genius, the appreciation of Einstein’s achievements only grows, until we get to the (possibly nonexistent) superstar who, now or in the next decade or so, sees a genuine “greatest blunder” buried in the general theory of relativity.

Einstein may be special—so well known through his writings in so many different spheres that the term “legend” is hardly appropriate. Here I want to tell a story and then make a statement about A E. Telling stories is something I do a lot, after more than thirty years of teaching physics.

Sometime around 1950 a mathematician friend at Princeton asked me if I would like to meet Einstein. At that time, I was a graduate student at Columbia University’s Nevis Laboratories, working on its new Synchrocyclotron. Then the most powerful particle accelerator in the world, the machine could accelerate protons to the incredible energy of 400 million electron volts (400 MeV).  For scale, the equivalent machine today at Fermilab reaches 2,000 billion electron volts (2 TeV). And so it happened that my best friend from high school, Martin Klein—then a graduate student in theoretical physics at MIT—and I were seated on a bench in Princeton waiting for the Master to pass by with his assistant, Ernst Strauss, who had arranged an introduction. My more-than-fifty-year-old recollection is shaky and would not hold up in any court, but here’s how I remember it:

Sure enough, here they come. Einstein has on his usual costume — sweatshirt, baggy pants, sandals. They stop, and Ernst asks him if he would mind meeting some physics graduate students. “No, it will be a pleasure,” says Einstein.

We stand, and he asks Martin, “What are you working on?”
 
“Quantum theory,” says Martin. 

“Ach!  A waste of time!” Einstein then turns to me, and I hasten to say that I am doing experimental research on the properties of pions. These subnuclear particles were discovered a few years earlier in cosmic rays and were supposed to produce the strong force that hold the atomic nucleus together; the Nevis accelerator was a prolific source of them.

Einstein nods, then shakes his head and says something to the effect that it is already impossible to explain the existence of the electron so why spend so much effort on these newer particles? He bids us a cheery goodbye, having crushed us both in about thirty seconds. However, we were way up in the clouds. We had met and talked physics with Einstein! The thrill was unimaginable—what he said hadn’t mattered at all. Since then, Martin has become a leading scholar in the history of physics and a coeditor of Einstein’s papers, and I helped to discover additional useless fundamental particles, like neutrinos and quarks.

Why was I not upset by the meeting with Einstein?

This question involves how physicists evaluate major physics achievements, which is clearly different from how laypeople, even science groupies, evaluate them. If we consider a particular discovery or creation—for example, the general theory of relativity—then the appreciation of this seminal achievement will still be driven by history and personality.  Physicists recognize that the general theory was uniquely Einstein’s. He labored over it for a decade. His drive was not to explain a plethora of experimental results but to express the beauty and simplicity of nature. (His personification for nature was Der Alte, the old one.) 

Experiments were of course relevant, and over the decades after the 1916 paper, experiments of awesome precision affirmed that relativity might be a correct theory of gravitation.

So, was that lonely mind influenced? Yes, by Ernst Mach, by James Clerk Maxwell, by mathematical helpers—but in this search for a more profound simplicity in the nature of space, time, and gravity he was very much alone. 

Let me place myself somewhere on the bell curve of physicists—say, midpoint—and try to describe how physicists think about Einstein and the very few others who have made major breakthroughs: Newton, Maxwell, Bohr, Schrödinger, Heisenberg, Dirac, and Einstein. Every one of us has such a list, and my guess is that these names would be included in most. But to me, Newton and Einstein, in true Christmas-tree fashion, flash on and off. They were all alone in what they did. Yes, they had guys nearby: Henri Poincaré, Hendrik Lorentz, and Mach for Einstein; Robert Hooke and Gottfried von Leibniz for Newton, but these two were truly far out there, all alone.

The Einstein prejudice, for me, stemmed from my reading, aged about sixteen, of The Evolution of Physics, a popularization for nonscientists coauthored by Einstein and the Polish physicist Leopold Infeld. The book introduced the theory of relativity but also provided an insight into Einstein’s philosophy. What I recall most vividly was its opening metaphor: The authors compared science to a detective story. The way I tell it now, there is a white Ford, a barking dog, a bloody glove, of course a body or two. These and other clues are meticulously recorded and ultimately the detective (scientist) assembles the suspects and solves the crime, thereby accounting for all the clues.

Here I should record my subjective reaction to other major physics breakthroughs. Somewhere in high school—before 1939—I  read about Niels Bohr’s use of the concept of quantum energy levels in the structure of the hydrogen atom. Bohr blended a mixture of classical physics and his ad-hoc and shocking introduction of discreteness in atomic structure. He also adopted the Planck-Einstein concept of photons—bundles of light energy. The precise wavelengths (colors) of the many spectral lines of the hydrogen atom followed, after some lines of simple algebra. What made teen-aged Leon gasp with excitement was the collection of symbols clustering in front of the terms that enumerated the spectral lines. There one found the velocity of light, the charge on the electron, Planck’s constant, and assorted two’s and pi’s.

How could these constants, originating in totally different contexts, come into a description of the hydrogen atom and correctly and precisely give rise to the spectral lines emerging from glowing hydrogen gas?  I recall putting the book down and pacing our house, frustrated that there was no one with whom I could converse about this amazing discovery.  I learned an incredibly profound concept about physics: that an idea articulated and composed in the music of mathematics can precisely describe a complex but beautiful piece of nature.

Another graphic example of creative imagination and a profound sense of the respect that nature has for mathematics is Paul Dirac’s famous equation describing the electron. Dirac was obsessed by the beauty of equations; his equation for the electron was not only beautiful but also unexpectedly fruitful. In the sense that the square root of four is plus two but also minus two, the equation for the electron predicted two electron-type particles: a negative electron (Dirac’s objective) but also a positive electron. Dirac’s urge to elegance and beauty had uncovered a revolution in physics: the existence of antimatter. For every particle—electrons, protons, neutrinos, quarks—there must be an antiparticle. What Dirac’s epiphany illustrates is the deep influence of the concept of symmetry on the physics of the twentieth century. Since symmetry thrives in mathematics, in arts and architecture, in music and mathematics, its influence in physics not only sparked a revolution in theoretical science but also acted as a unifying connection to the humanities.

Now comes Einstein’s year of glory.

It has been pointed out in many places that Einstein’s miracle year of 1905 followed several years of discouragement—first with the entire process of being examined for his PhD degree, then with the slow acceptance of his thesis paper, and finally with the need for and the difficulty of finding a job in his chosen field. Sitting as a clerk in the federal patent office in Bern, Einstein—then twenty-six years old—caught fire and in five stunning papers, all published in 1905, the kid solved three of the most important problems in the physics of his time: the existence and reality of atoms and molecules, the quantum behavior of photons, and a new statement of the principle of inertia, first enunciated by Galileo some three hundred years earlier. Since inertia and relativity are closely connected concepts, the new statement is now referred to as Einstein’s special theory of relativity.

By the time I was in college, Einstein’s renown was so strong that it had to color my judgment as to the depth of his paper on special relativity. But my student days were obsessed with questions: Where did he get this idea?  Why Einstein?  How could such a simple statement of a concept or a principle have such profound implications?

Einstein was examining patents during the long days and presumably working on physics nights and weekends—why? He had not been driven by some experimental breakthrough (although there were growing experimentally inspired doubts about the Newtonian worldview) but by an aesthetic and deep physical sense of the accordance of symmetry with nature. Since symmetry is closely associated with beauty and simplicity, we come easily to a belief in Einstein’s view of how nature works.

The key word, which we learn in graduate courses but which should be taught in high school science courses, is “invariance.” When a physical system is observed from different points of view, or when the system is subject to tortures that only physicists are capable of imagining, it is of intense interest to see what changes and what doesn’t change. Does part of the system change? The total energy? The entire system? If nothing changes, the system is invariant.  This is nature at its simplest. The system’s laws of physics do not care whether observer Joe studies the system while at rest (that is, with the same velocity) or whether Moe, equally adept, speeds by with a huge relative velocity. Moe, the careful physicist, sees Joe and all of Joe’s experiments from his (Moe’s) viewpoint as he moves past, but Moe also sees the same laws—the same rules. This is true, says Einstein, no matter what the relative velocity. Stated in textbook style, the laws of physics are the same for all observers moving with constant velocity. 

This was not a departure from Newtonian science, but Einstein was now also dealing with the phenomena of electricity and magnetism. Maxwell had summarized those experimental laws brilliantly in 1860. The summary of the relevant experiments led to Maxwell’s discovery that light was an electromagnetic phenomenon. Combined electric and magnetic forces, vibrating and escaping from their wires into space, traveled at the magnificent speed of 186,000 miles per second. The velocity of light, said Einstein, was a law of physics, the same for all observers!  Only in this way could the invariance both of Newtonian systems and Maxwellian systems be respected.  So simple! But so profound. The assertions, taken together, constitute the special theory of relativity and therefore a revolution in our concepts of space, time, and energy.

All the confusions and desperate efforts to understand the experiments that were crowding classical physics were swept away by these statements. Who could not love the iconoclast who blew up something called “the luminiferous aether”?  Could any of the great scientists of the new century—Poincaré, Lorentz—could any of them have created this idea?  Have some fun: Raise that issue at the faculty club near the physicists’ table and then avoid the flying debris and lurid language as the physicists’ table erupts.

The special theory combines the two ideas: The velocity of light is the same  (invariant) for all observers, and the laws of physics are the same (invariant) for all observers moving at constant velocity. The symmetry and elegance of electromagnetism is thereby preserved—but when these ideas were applied to Newton’s mechanics, the world changed. This is Einstein’s special theory, and the consequences—economic, technological, and scientific—were as profound as the statement was simple.

The startling thing about special relativity is the engineering applications.
We should note that nuclear energy itself is not a consequence of the theory, but there are a plethora of devices that make use of one of the major predictions: that as particles move at velocities approaching the speed of light, there is an increase of mass. Devices whose designs depend upon this effect are large radio frequency amplifiers (klystrons); electron accelerators, used by the thousands in cancer therapy; electron microscopes; high-voltage television tubes; industrial accelerators for sterilization and for control of manufacturing processes, such as thickness measurements; and, most spectacularly, high-energy particle accelerators, which advance our knowledge of the structure of matter and energy.

Another ever-increasing application is the use of high-energy beams of electrons to produce “synchrotron light,” an intense source of X rays used to etch silicon elements for microelectronics and to give chemists and biologists graphic photographs of the three-dimensional molecular structure of new materials, new chemicals, and data on DNA and other biological structures. All this from a patent clerk with an attitude.

Although the accumulated contribution of these devices to the GNP is hundreds of billions of dollars, that all pales into insignificance compared with the revolutionary impact of Einstein’s conceptual breakthrough. Much of this hovers around the new, subjective interpretation of time, and it is here that most of us plebian professors and Nobel laureates can only shake our heads in wonder and gratitude.

When Moe, traveling at a high speed relative to Joe, records the same phenomena as Joe is recording, the numbers are of course different. Joe, for example, locates, say, an electron (a component of the system he is studying) at these coordinates: x = 6.2,  y = 9.6,  z = 27.3 (all in appropriate units—say, meters).  He gives as its velocity v = 9.6 x 108 m/s along his x-axis.  Moe, looking at the same dumb electron, will have different numbers, because his coordinates—his x’s, y’s, and z’s—will be different. The electron’s velocity measured in Moe’s lab will be different. If we designate positions and velocity as seen by Joe as x, y, z and v (along x) and t (time when the measurements were made), Joe’s electron coordinates are x, y, z, t. In Moe’s lab let’s call his measurements: x', y', z', v' and t'.

The laws of physics should not depend on the system or the observer, because there is no way to tell whether Joe or Moe or both are moving. We know only their velocity relative to the system. With a little algebra, we can find the relation of these two sets of coordinates. So far, Newton and all his progeny would be happy. However, Newton would immediately say that t' = t—that is, the clocks in Joe and Moe’s lab must read the same time intervals. But in special relativity, the rate of timekeeping may not be the same, and the discrepancy will increase as the relative velocities approach the velocity of light. The new and bizarre aspects of time are the fault of Einstein’s equations, which twist and embed time with space, to the despair of the earnest undergraduate.

In the hundred years since the special theory was proposed, this prediction—that clocks, synchronized when, for example, Joe and Moe are at relative rest, run at different rates when Moe revs up his lab and speeds off—has been borne out.

Another story: My PhD thesis experiment, carried out in 1950 (hardly a man is now alive….) used a natural clock, the radioactive particle called a muon. Accelerators produce muons at very high velocities, but one can also find muons essentially at rest. At rest, their characteristic lifetime—the length of time it takes some set fraction of your muons to decay—has been carefully measured. When the muons are moving at around 98 percent of the velocity of light, their lifetime is extended fivefold! Gee, if Moe could travel at that speed his lifetime would be about four hundred years!

The catch is, he would not be aware of all that extra longevity until he was able to visit his pal Joe and find that whereas only, say, ten years had passed as far as he was concerned, Joe was now fifty years older. Relative to Moe’s clock, Joe’s had speeded up fivefold. This is equivalent to Moe’s clock, relative to Joe, slowing down to allow him to live to age four hundred, as clocked by envious Joe.

This profound alteration in the nature of time is but one example of the deep philosophical implications of Einstein’s revelations about space and time, the pillars of the world we inhabit. It is for me unimaginable that this sweat-shirted shuffler, totally unappreciative of two such promising and handsome grad students, could have had the crystalline clarity of thought to see, discover, compose, invent so much simplicity and beauty in our world.


Excerpted from My Einstein by John Brockman, Copyright © 2006 by John Brockman. Published by Pantheon, a division of Random House, Inc.


 

 

 

John Brockman, Editor and Publisher
Russell Weinberger, Associate Publisher

contact: [email protected]
Copyright © 2006 By
Edge Foundation, Inc
All Rights Reserved.

|Top|