Skip to contentSkip to site navigation

Why Einstein?

By James Challey
Lecturer in Physics and Director of the Science, Technology and Society Program

When Albert Einstein died in the early morning of April 18, 1955, he was without question the best-known scientist in the world. The New York Times carried the news on its front page. In the Washington Post Herblock, perhaps the premier political cartoonist of his day, commemorated the event with a sketch of the cosmos, in which Earth is identified by a simple label, “Albert Einstein lived here.”1 Today, exactly 50 years after his death and exactly 100 years after the publication of the papers that would change the course of modern physics Einstein remains the most famous figure of 20th century science. A picture of Einstein is still instantly recognizable: the shock of white hair perpetually uncombed, the deep, gentle eyes, the bushy mustache, and the casual indifference to dress. In the popular media Einstein continues to serve as an icon of the scientist, at once a stereotype and an ideal type.

For the historian of science this presents an intriguing puzzle. Why Einstein? Persuasive arguments could be made that several other major scientists of the 20th century, such as Niels Bohr, John von Neumann, Linus Pauling or James Watson, have had more direct impact on our lives than Einstein has. That they remain relatively unknown suggests that the answer to Einstein’s enduring public image lies not in the consequences of his scientific work or its technological impact, but in a much more complex interplay between the content or the style of Einstein’s science and the social world around him. Indeed, the question might even be reversed; how did Einstein project himself and his work in such a way that it captured and shaped the public concept of what science should be?2

It is not surprising that Einstein has been a challenge for biographers, who tend to fall into two separate traditions, those who study Einstein the scientist and those who portray Einstein the man.3 Through the 1980’s most biographies were based on Einstein’s published work, the autobiographies of Einstein’s colleagues, and the personal recollections of the authors. Because of the unreliability of some sources and the tricks that memory plays on the past, the inevitable, though unintentional, result was the creation and perpetuation of an Einstein legend. Fortunately, in the past 15 years, the availability of Einstein’s private papers and letters, of which the Bergreen collection is an important part, has dramatically improved our understanding of Einstein the man and the scientist.4 In these documents Einstein has left, in his candid and graceful writing, fascinating glimpses into his scientific work and the social context in which it took place. In them we can now begin to see the seamless web that connected Einstein the scientist and Einstein the public figure.

The Making of a Physicist

One of the first fruits of the publication of the complete Einstein papers has been the demolition of many of the legends of Einstein’s childhood. Born March 14, 1879 in Ulm, Germany, Einstein was the only son of a thoroughly secular Jewish manufacturer of electrical equipment. Contrary to conventional accounts, Einstein had no difficulty talking as a toddler and did very well in school, regularly winning prizes in mathematics and Latin. The school experience, however, was unpleasant; he disliked the regimentation within the classroom, and he was a loner among his classmates. Accustomed to being mothered, the young Einstein was very shy among other boys, had no interest in games or sports, and by his early teens was aware that he was a Jew in an overwhelmingly Christian society.5

Einstein’s real education took place outside the school, in a habit of wide reading that he kept his entire life and in conversations with older relatives and his father’s business associates. Despite the popular image of the solitary thinker, Einstein always needed a sounding board, someone with whom he could talk through his ideas.6 When the family moved to Italy in search of better business prospects, Einstein was left behind in Munich to complete his education. It soon became unbearable, and at age fifteen he quit school, rejoined his family and moped about for a year, finally deciding to study electrical engineering at a prestigious polytechnical institute in Zurich. The entrance examinations at the Eidgenossische Technische Hochschule (ETH) were quite rigorous, and Einstein failed the sections on biology, chemistry and French. But his scores in physics and mathematics were so strong he was admitted, pending a remedial year in a Swiss boarding school.

Finally, in 1896, Einstein enrolled at the ETH and almost immediately blossomed, personally and intellectually. He found a close circle of friends, with whom he spent long hours playing music and discussing science, philosophy and literature.

There were altogether only two examinations; aside from these, one could just about do as one pleased. This was especially the case if one had a friend [Marcel Grossman], as did I, who attended the lectures regularly and who worked over their content conscientiously. This gave one freedom in the choice of pursuits until a few months before the examination, a freedom which I enjoyed to a great extent…7

In this setting the shy adolescent became a self-confident, almost cocky, young man. Nonchalant about attending classes, Einstein read widely and deeply in the literature of physics and mathematics, mastering the classic texts and becoming completely familiar with the central problems of late 19th century physics. But there were still the exams to face. Thanks to Grossman’s meticulous notes and several weeks of intense cramming, Einstein posted a set of very respectable exam scores and graduated in 1900.

There followed two turbulent years, during which Einstein took a series of part-time teaching jobs and looked for a permanent position. During this time Einstein’s passionate love affair with Mileva Maric, begun when they were both ETH students, resulted in the birth of a daughter.8 Finally, thanks in part to the influence of Marcel Grossman’s father, Einstein obtained a position as patent examiner in the Swiss Patent Office in Bern. It was nearly the perfect job for the ambitious young physicist. The position paid a modest salary, but enough to allow Einstein to marry Mileva and settle down, and the actual work of examining patent applications, which Einstein enjoyed, took only a few hours each day. That left plenty of time to think about physics.9

Einstein the Physicist

During his first two years at the Patent Office Einstein published four short papers on thermodynamics and statistical mechanics in the Annalen der Physik, the leading German physics journal. But these articles, the first careful steps of a promising young physicist, gave no warning of Einstein’s outburst of scientific creativity in 1905. In one year, in a display of virtuosity that can only be compared to Newton’s annus mirabilis, Einstein published five papers, each one of which would have made his reputation in physics.10 Two of them extended statistical mechanics into new realms; the other three created modern physics.

One of the problems Einstein began thinking about was blackbody (or thermal) radiation, the energy emitted by heated objects. The problem was of particular interest for two reasons. First, the explanation of blackbody radiation was expected to lie in the combination of exactly the three theories that had been the greatest triumphs of 19th century classical physics. In thermodynamics heat had been explained as a form of motion; the theory of electromagnetism had explained all of electricity and magnetism on the basis of the motion of electric charges, and in optics light was understood as an electromagnetic wave. The second reason for interest was that the problem had stubbornly resisted solution. In 1900 Max Planck, the Professor of Physics at the University of Berlin and one of the world’s leading theoretical physicists, had managed to derive an equation correctly describing blackbody radiation, but at an embarrassing price. He had to assume that the energy of the light waves could exist only in certain sizes, integral multiples of a particular unit of energy he called an “energy quantum.” The assumption was just as bizarre as saying that water waves can exist only in certain, fixed heights (1 foot, 2 feet, 3 feet, etc) but not at any intermediate heights. So Planck regarded his quantum hypothesis as completely artificial, a kind of bookkeeping trick that could be removed in the final form of the theory. But by 1905, despite the repeated efforts of Planck and of many other leading physicists, the problem remained, more troubling than ever.11

On March 17, 1905, Einstein submitted a paper to the leading German physics journal entitled, “On a heuristic point of view concerning the generation and conversion of light.”12 After a careful review and critique of Planck’s work, Einstein made a bold proposal: the energy quanta are real. Light behaves as if it consists, not of waves, but of a stream of independent bundles of energy he called light-quanta. He then went on to describe an experiment on a little-known phenomenon called the photoelectric effect that could test his proposal and that could directly measure the energy of the light particles.

The paper was breathtaking in its audacity; indeed, it was his only scientific work that Einstein himself called revolutionary.13 In the first place, it flew in the face of everything that had been learned over the previous century about the wave properties of light. The electromagnetic wave theory of light and such well-known phenomena as the interference and diffraction of light simply could not be explained by a theory in which light consisted of independent particles, To make matters worse, the light quanta had a strange combination of wave and particle properties. Each light quantum contained a specific chunk of energy, but the amount was determined by the wavelength of the light, a wave property. (This duality in the nature of light would haunt Einstein for the rest of his life. In 1951 he wrote to his old friend Besso, “ All these fifty years of pondering have not brought me closer to answering the question, What are light quanta?”14) And finally, the experiment Einstein had described to test his suggestion turned out to be technically very difficult to do. So it is not surprising that Einstein’s paper on the photoelectric effect generated almost no response for nearly a decade.

Things were quite different for the paper submitted to the Annalen der Physik on June 30. The title of the paper, “On the Electrodynamics of moving Bodies,” would have caught the attention of nearly every reader, for it succinctly defined one of the major theoretical challenges facing classical physics at the end of the 19th century.15 The problem, which had resisted the efforts of such major figures as Poincaré and Lorentz, had evolved into a paradox. On one hand, there was the principle known as Galilean relativity, which held that the laws of physics must be the same in any inertial reference frame, i.e., any setting which is moving at a constant speed in a straight line. To give an example of which Einstein was fond, an experimenter working in a closed railway car moving down a straight track at a steady speed will get exactly the same results for any experiment as another experimenter working on the station platform. In fact, each experimenter could reasonably claim to be at ‘”rest” and the other one in motion. As a consequence each could translate their results into the other’s framework simply by including the relative motion. In the train example an object thrown forward at 20 mph by the experimenter on a train going 30 mph is going 50 mph when measured by the observer on the platform. During the 19th century the principle of Galilean or classical relativity seemed so obvious, so ordinary, that most textbooks didn’t bother to state it explicitly.

The other half of the paradox involved the speed of light. The same set of Maxwell’s equations that explained the nature of light as a wave consisting of interacting electrical and magnetic fields also predicted the speed of light. This speed depends only on the electrical and magnetic properties of the medium, but not on the motion of the source or of the observer. But this contradicts the principle of classical relativity, which asserts that the light beams produced by two experimenters in relative motion should travel at different speeds. By 1905 there had been numerous attempts to resolve the problem by experiment, but despite some extremely sophisticated methods no change in the speed of light due to inertial motion could be detected. On the theoretical side, several attempts to modify Maxwell’s equations for moving objects proved unsatisfactory. What had begun as an intriguing anomaly had become a crisis that threatened the foundations of physics.16 Either the principle of classical relativity was wrong and the laws of physics are different for moving objects, or Maxwell’s theory of electromagnetism and light was wrong.

In his paper Einstein’s masterstroke, at once bold and elegant, was to accept the two halves of the contradiction and make them the postulates of a new theory of relativity, the special theory of relativity. The paper begins with the statement of the two postulates: I, that laws of physics are the same in all inertial reference frames; and II, that the speed of light is a constant for all sources and observers. Einstein then examined what had to be true for these two statements to be consistent. The concept linking them is velocity, the ratio of distance and time. In a question of breathtaking simplicity and depth, Einstein asked: How do experimenters in different reference frames measure distance and time? In the first of a series of elegant thought experiments he began by showing that events that one experimenter regards as simultaneous will be observed occurring at different times by someone in a different reference frame. He then proceeded to show that, in addition, an observer in a given inertial reference frame will see moving lengths contract, the masses of a moving objects increase, and moving clocks run slow. Then he was able to show that these new concepts of space, time and mass led to new rules for translating the measurements in one reference frame into another (the “Lorentz transformations”). Finally Einstein could show that with these transformations Maxwell’s equations were true in any reference frame. A few weeks later Einstein realized that the new theory of special relativity also implied that matter and energy should be equivalent and interchangeable. In a short paper17 that appeared late in 1905, Einstein proved that E = mc2.

The two 1905 papers on special relativity attracted immediate attention. At the University of Berlin Max Planck included special relativity in his physics seminar for 1905-06. One of his students, Max von Laue, published several papers working out further implications of Einstein’s work. In Zurich Hermann Minkowski, one of Einstein’s former mathematics professors, reformulated the new concepts of space and time into the idea of a four-dimensional “spacetime.” Propelled by the importance and originality of his work and guided by powerful patrons such as Planck, Einstein began a rapid rise through the notoriously hierarchical German academic system. After a series of brief teaching appointments in Bern, Zurich and Prague, in 1913 Einstein was elected to the Prussian Academy of Sciences and appointed to a salaried, non-teaching research position in Berlin. In his nomination speech to the Prussian Academy Planck captured the sense of excitement Einstein had generated.18

Summing up, we may say that there is hardly one among the great problems in which modern physics is so rich, to which Einstein has not made an important contribution. That he may sometimes have missed the target in his speculations, as for example, in his hypothesis of light quanta, cannot really be held too much against him, for it is not possible to introduce fundamentally new ideas, even in the most exact sciences, without occasionally taking a risk.

In the meantime, Einstein was trying to extend relativity beyond inertial reference frames to include situations in which objects are accelerated. The breakthrough came in 1907 in what Einstein later called “the happiest thought of my life.”19

I was sitting in a chair in the patent office in Berne when all of a sudden a thought occurred to me: If a person falls freely, he will not feel his own weight. I was startled. This simple thought made a deep impression on me. It impelled me toward a theory of gravitation.

The awareness that being in a uniformly accelerated reference frame (falling freely) is fully equivalent to being in a gravitational field later became known as the principle of equivalence.20 From this deceptively simple observation Einstein began to develop the general theory of relativity, which he finally completed in 1915.21 In these works Einstein laid the foundations of a general theory of gravitation in which gravity is explained as the curvature of spacetime created by matter. With his characteristic instinct for the concrete example, Einstein went on to deduce two consequences of the theory that could be observed. One of these was well-known, the slow shift in the orbit of Mercury that had resisted explanation in terms of Newton’s theory of gravity, but which Einstein could explain in terms of the intense gravitational field created by the sun. The second was entirely new; light from distant stars passing through the curved spacetime near the surface of the sun should follow a curved path. The deviation from the straight line path followed in empty space would be small, but it should, Einstein calculated, be large enough to be observed during a total eclipse of the sun.

While working on gravity and the general theory of relativity Einstein’s attention was brought back to the 1905 hypothesis of the light quantum by two events that would have far reaching consequences. The first was the publication in 1914 of the results of experiments undertaken by Robert Millikan, a young American physicist working in Chicago who set out to test Einstein’s predictions on the photoelectric effect. It took Millikan over five years to complete the difficult experiments, but the results matched Einstein’s predictions almost perfectly. Despite this clear confirmation of Einstein’s light quantum hypothesis, the concept of light as independent particles remained so foreign to the standard theory of optics that most physicists, including Millikan himself, continued to doubt that light quanta could really exist.22 The second major development in the quantum theory of light was Niels Bohr’s theory of the atom, laid out in a series of papers beginning in 1913.23 In Bohr’s theory the atom consisted of a dense nucleus surrounded by electrons which circled the nucleus only in orbits corresponding to certain, quantized energy levels. When an electron “jumped” from a higher to a lower energy level, it emitted the difference in energy in the form of a light quantum. Using his calculations, first for the hydrogen atom and then for more complex atoms, Bohr was able to show the predicted values for the light emitted matched very closely the known wavelengths of the light emitted by atoms of the different elements. Bohr’s work was an immediate sensation and set in motion the development of modern atomic theory and the elaboration of quantum mechanics. But, ironically, it was also the beginning of the development of the two profoundly different approaches to quantum mechanics that would separate Bohr and Einstein, ultimately leaving Einstein isolated from the physics community.

The World War I years were spent in Berlin. Although the deeply pacifist Einstein privately viewed Germany’s role in the conflict with revulsion, he focused his energies on his research. In addition to completing the first full version of the general theory of relativity, Einstein did further work on the properties of light quanta and began exploring the implications of the general theory for cosmology. While virtually unknown outside the scientific community, Einstein had already established a reputation as one of the major figures in physics, a key member of an elite circle that included the giants of the older generation such as Planck and Lorentz and the rising stars of the new generation such as von Laue, Bohr and soon Louis de Broglie. Beginning in 1912 Einstein was regularly nominated for the Nobel Prize.

In 1919 a team of British astronomers led by Sir Arthur Eddington traveled to the South Atlantic to be present for a total eclipse of the sun. During the few minutes of total eclipse they photographed the background stars visible near the edge of the darkened sun. Returning to London Eddington and his team carefully compared the apparent positions of the stars near the sun with their positions of the same stars as seen without the sun present. The results, announced by Eddington before a joint meeting of the Royal Society of London and the Royal Astronomical Society, confirmed Einstein’s prediction.24 Although this was an important test of general relativity, it was nearly the only direct test possible at the time. So, by itself Eddington’s observation of the bending of light did little to convince the skeptical and the cautious, which included the Nobel Prize award committee.

Notoriously conservative lest they make an award for work later found to be incorrect, the Nobel committee maintained that the evidence for general relativity was still insufficient. Meanwhile, each year, Einstein was being nominated for the physics prize, with growing insistence and impatience, by virtually all the major physicists. Finally, in 1922, Einstein was awarded the 1921 Nobel Prize for physics. In the citation the Nobel committee pointedly avoided any mention of relativity. Instead Einstein was honored “for his services to theoretical physics, and especially for his discovery of the law of the photoelectric effect.”25 That Einstein’s most radical theory was honored by its most conservative institution remains one of the great ironies of modern science.

As is so often the case, the Nobel Prize merely confirmed the consensus of the physics community. By the early 1920’s, with the restrictions of the war lifted, Einstein was in great demand on the scientific lecture circuit. The news of the Nobel reached Einstein aboard a ship bound for Japan, where he gave a series of lectures on physics, as he had in several countries along the way. The previous year he had given major lectures in France, England and the United States. The following year he did an extensive lecture tour in South America. Although Einstein continued to work on cosmological theory, he was increasingly drawn into an extended dialogue with Niels Bohr on the meaning of the stunning new developments transforming quantum theory. In the hands of Bohr and a group of brilliant young protégés including Werner Heisenberg and Max Born the light quanta Einstein had proposed became photons, particles of energy that yet retain wave properties. The work of Louis de Broglie extended the wave-particle duality to matter. The explanation for the behavior of the atom was being written in the language of probability, and the Uncertainty Principle suggested that there were real limits to what science could know.26 It was a concept of physical theory, and of science itself, that Einstein could not accept. Einstein deeply shared the 19th century, and ultimately Enlightenment, belief that the universe was governed by universal laws that humans could discover through rational enquiry. The existence of a wave-particle duality was, for him, simple evidence that a deeper understanding must be sought. As for the role of chance, Einstein stated his position with a moving, and much quoted, eloquence in a letter to Max Born in 1926:27

Quantum mechanics is very impressive. But an inner voice tells me that it is not yet the real thing. The theory produces a good deal but hardly brings us closer to the secret of the Old One. I am at all events convinced that He does not play dice.”

In 1927 the Solvay Conference in Brussels brought together all the participants in the revolutions that had created modern physics. The real drama occurred outside the formal sessions, in a debate extending over several days in which Einstein raised objections, pointed out logical inconsistencies and invented ingenious thought experiments to undermine the validity of the new quantum theory. And every day, with equal conviction, intensity and personal respect, Niels Bohr answered Einstein, point-by-point.28

It was the last major scientific conference Einstein attended. Increasingly regarded as outside the mainstream of physics, Einstein continued working on general relativity. But more and more of his effort went into trying to formulate a unified field theory, a theory that would reduce the two great forces of nature, gravity and electromagnetism, to a single origin and explanation. Einstein was quite aware that the new generation of physicists considered his efforts irrelevant. But he persisted for over twenty-five more years. Among the papers at his bedside when he died was yet another draft of a unified field theory.29

Einstein as a Public Figure

Prior to 1919 Einstein was virtually unknown outside the scientific community. German newspapers carried the news of his appointment to the prestigious Prussian Academy and a few notices of lectures and publications, but elsewhere he was completely unknown.30 Scientific American first mentioned Einstein’s 1905 work on special relativity only in 1911. The following year an article in the same magazine concluded that there was not enough experimental evidence to decide if the theory were true.31

Just now, however, the ‘Principle of Relativity’ seems to be irresistibly fascinating to mathematicians, but equally abhorrent to that host of physicists who can no more conceive of time as a function of velocity than they can imagine space to be curved or picture for themselves a fourth dimension.

At the same time, it should be noted that the only scientist in the period before World War I to be widely covered in the popular press was Wilhelm Roentgen, whose discovery of x-rays quickly led to medical applications and widespread public interest.32

Einstein became a public figure in 1919, suddenly and dramatically. If, as one of his biographers has argued, Einstein was a creation of the media, then some clues to the nature of the Einstein image can be found in the details of the creation process.33 On Nov. 7, 1919 the Times of London carried a story that described in considerable detail the joint meeting held the day before between the Royal Society of London and the Royal Astronomical Society. The purpose of the meeting was the presentation by Sir Frank Dyson, the Astronomer Royal, of the results of the measurement of the bending of starlight made by the Eddington expedition. Under headlines that proclaimed “Revolution in Science. New Theory of the Universe. Newtonian Ideas Overthrown,” ,the article summarized the proceedings:34

It was generally accepted that the observations were decisive in the verifying of the prediction of the famous physicist, Einstein, stated by the President of the Royal Society as being the most remarkable scientific event since the discovery of the predicted existence of the planet Neptune.

Despite the finality suggested by the headline, the article went on to point out that there was in the discussion following the presentation of the data some uncertainty about both the observations and what they meant for Einstein’s theory.

Even the President of the Royal Society, in stating that they had just listened to “one of the most momentous, if not the most momentous, pronouncements of human thought,” had to confess that no one had yet succeeded in stating in clear language what the theory of Einstein really was. It was accepted, however, that Einstein, on the basis of his theory, had made thee predictions. The first, as to the motion of the planet Mercury, had been verified. The second, as to the existence and the degree of deflection of light as it passed the sphere of influence of the sun, had now been verified. As to the third, which depended on spectroscopic observations there was still uncertainty. But he was confident that the Einstein theory must now be reckoned with, and that our conceptions of the fabric of the universe must be fundamentally altered.

In its final paragraph the article indicates that one of these new conceptions is that Newtonian absolute space must be replaced by the idea that space can be curved or “warped.” The article concludes with a remarkably judicious appraisal of the logical status of Einstein’s general theory of relativity.

His predictions in two of three cases have now been verified, but he question remains open as to whether the verifications prove the theory from which the predictions were verified.

The London Times published several follow-up articles over the next few days, including a short article describing the special and general theories written by Einstein, and papers in Holland and in Germany also picked up the story.35 But it was the New York Times that contributed most to the Einstein mystique, with a story introduced by a breathless 6-part headline.39

Lights All Askew in the Heavens
Men of Science More or Less Agog Over Results of Eclipse Observations.
Einstein Theory Triumphs
Stars Not Where They Seemed or Were Calculated to be, but Nobody Need Worry.
A Book for 12 Wise Men
No More in All the World Could Comprehend it, Said Einstein When His Daring Publishers Accepted It.

Over the next several weeks Einstein was mentioned frequently in articles and editorial remarks. “This news is distinctly shocking and apprehensions for the safety of confidence even in the multiplication table will arise.” And if Einstein’s suggestion that space is curved and finite, the Times felt that scientists “are under obligation to tell us what lies beyond it.”37 In the case of the Times this was the beginning of a close coverage in which, for the rest of his life, articles about, and occasionally by, Einstein appear at least once each year.38

The effect of this publicity became apparent sixteen months later when Einstein made his first visit to the United States. The main purpose of the trip, organized and led by Chaim Weizmann, the leader of the World Zionist Organization, was to raise money for the Zionist cause. In this effort Einstein’s role was to speak on behalf of the Hebrew University then being planned. But, no doubt to Weizmann’s chagrin, the public spotlight was on Einstein and relativity. In New York Einstein was met by cheering crowds lining the streets as his car slowly made its way from the pier to his hotel. The New York Times covered the visit in detail, starting with a front-page story describing his arrival.37

Prof. Einstein Here, Explains Relativity
“Poet in Science,” Says It is a Theory of Space and Time, But It Baffles Reporters.
Seeks Aid for Palestine
Thousands Wait Four Hours to Welcome Theorist and His Party to America

Interviewed as he waited for the ship to dock, Einstein was described as wearing “a faded gray raincoat and a flopping black felt hat that nearly concealed the gray hair that straggled over his ears,” holding a pipe in one hand and his violin in the other.40

But underneath his shaggy locks was a scientific mind whose deductions have staggered the ablest intellects of Europe. One of his traveling companions described him as an “intuitive physicist” whose speculative imagination is so vast that it senses great natural laws long before the reasoning faculty grasps and defines them.

At his first public appearance eight thousand people crowded inside, while another three thousand waited outside. Over the next two months large, enthusiastic crowds greeted Einstein as the entourage traveled to Washington, DC, Chicago and Boston. At the same time Einstein gave physics lectures at several major universities, including Columbia, City College, Chicago and Princeton.

By the time Einstein made his second visit to the United States eight years later, he had become a genuine celebrity. In New York he met with the Mayor and with John D. Rockefeller. His arrival in California was described as “one part show business, one part hero worship, and one part genuine affection.”41 Among other well-publicized events Einstein dined with Charlie Chaplin and William Randolph Hearst and then was Chaplin’s guest at the premiere of City Lights. Will Rogers quipped that Einstein met with everybody. “In fact, he made himself such a good fellow that nobody had the nerve to ask what his theory was.”42 The Einstein image had closed around itself; now he was famous for being famous.

Why Einstein?

Abraham Pais, one of the best of Einstein’s biographers, has suggested that there were a number of factors involved in the creation of the Einstein legend. One of these was an accident of timing; Einstein’s work became popular just in the wake of the horrors of World War I. In this context Einstein appears as a new figure, “carrying the message of a new order in the universe.”43 Although this claim is difficult to establish, it is certainly plausible. One of the issues that Einstein commented on in his first public articles and lectures was the need to overcome the divisions the War had created in the international scientific community.44 Another reason for Einstein’s public appeal, Pais argues, is the sense of mystery and wonder that surrounded his work, particularly on relativity. The main ideas, such as the bending of space, could not only be stated in simple, ordinary language, but they also lent themselves to vivid visual images. At one point Einstein himself suggested that “it is the mystery of non-understanding that appeals to [the general public].”45 In considerable measure this sense of wonder was the result of Einstein’s own descriptions of his work, and Pais maintains that Einstein’s abilities with the German language are second only to his abilities as a physicist. Finally, Pais argues that, at least until very late in life, when he was genuinely photogenic and usually enjoyed posing for artists and photographers, Einstein’s personal appearance played little role.46 (The sweatshirts and baggy trousers came only in retirement, and the one quirk most often commented upon by his friends, an aversion to wearing socks, is probably best explained by Einstein’s allergy to wool.)

Although the quantity and quality of Einstein’s work declined just as his public activity and image increased, Einstein’s public image rested, first and foremost, on his role as a scientist. Undoubtedly part of the explanation for the Einstein’s fame is that he fulfilled, and indeed came to embody, the public conception of what science was. The main features of that conception can be seen in the newspaper accounts of Einstein. Science is abstract conceptual thought, the highest and most difficult form of thinking. Scientific talent is rare and inexplicable, pure genus. Science is above such mundane activities and concerns as dress and social convention. Even the reference to the limited number of people who could understand his work was a common trope, going back at least to Newton. Finally, science was seen as an enterprise apart from its applications. Even after the nuclear bombing of Hiroshima and Nagasaki, Einstein and his work represented a purity, an ideal, that contrasted with selfish goals and technological applications. Thus it is no surprise that Einstein, who rejected the concept of a personal God, was so often asked to compare science and religion. In short, one important reason for the Einstein legend is due, less to the content of his science, but to the way he personified the public image of what a scientist should be.

Recent work in the history, philosophy and sociology of science has done much to challenge this conception of science and to point out how deeply Big Science is intertwined with its applications and how extensively military needs have shaped the directions of research. At the same time, the Einstein correspondence, now becoming available, is showing that, if the Einstein legend was a creation of the media, Einstein was no mere passive subject in the process. What is striking about many of the letters in the Bergreen Collection is Einstein’s clear awareness of his public image and his commitment and courage in using that image to achieve his goals of peace and social justice.

————

  1. Herblock’s cartoon is reproduced as the frontispiece in Abraham Pais, Einstein Lived Here (New York: Oxford University Press, 1994).
  2. One of the major themes in recent work in the field of science and technology studies has been to treat science as a form of rhetoric and to examine how and for whom particular scientific achievements become persuasive. See, for example, Bruno Latour, Science in Action (Cambridge: Harvard University Press, 1987).
  3. For Einstein’s scientific biography, see Abraham Pais, “Subtle is the Lord…”: The Science and the Life of Albert Einstein, (Oxford: Oxford University Press, 1982). The most recent full biography is Denis Brian, Einstein: A Life, (New York: John Wiley & Sons, 1996).
  4. Albert Einstein, The Collected Papers of Albert Einstein, edited by John Stachel et al and translated by Anna Beck, (Princeton: Princeton University Press, 1987 - ). The most recent volume covers the mid-1920s.
  5. Once settled in the United States, Einstein took up sailing, probably because it was a solitary activity. For someone who couldn’t swim, he was not a very good sailor and more than once had to be rescued on Long Island Sound. Brian, op.cit., pp. 1-8, 262-3.
  6. Some of Einstein’s most extended correspondence represents an attempt to carry on these conversations by mail, sometimes over a period of decades. Albert Einstein, Letters to Solovine, 1906 - 1955, (New York: Citadel Press, 1993), and Albert Einstein, Correspondence avec Michelle Besso, 1903 - 1955, (Paris: Hermann, 1979).
  7. Albert Einstein, “Autobiographical Notes,” translated by Paul Arthur Schilpp, in Paul Arthur Schilpp, ed., Albert Einstein: Philosopher-Scientist, (New York: Harper & Row, 1959), Vol. I, p. 17.
  8. Roger Highfield and Paul Carter, The Private Lives of Albert Einstein, (New York: St. Martin’s Press, 1994), chapters 4 and 5.
  9. For an excellent account of Einstein’s work in the Patent Office and how it influenced his early thinking about relativity, see Peter Galison, Einstein’s Clocks, Poincare’s Maps: Empires of Time, (New York: Norton, 2003).
  10. A very helpful bibliography of Einstein’s published work can be found in Schilpp, op.cit., Vol. II, pp. 694-760.
  11. For a history of the general problem of blackbody radiation, see Thomas S. Kuhn, Black-Body Theory and the Quantum Discontinuity, 1894 - 1912, (New York: Oxford Universtiy Press, 1978). The best biography of Planck is J.L. Heilbron, The Dilemmas of an Upright Man: Max Planck and the Fortunes of German Science, (Cambridge: Harvard University Press, 2000).
  12. A. Einstein, “Über einen die Erzeugung und Verwandlung des Lichtes betreffenden heuristischen Gesichtspunkt,” Annalen der Physik, ser. 4, vol. 17, pp. 132-148.
  13. Pais, ‘Subtle is the Lord…’, p. 376.
  14. Albert Einstein, Correspondence avec Michelle Besso, 1903 - 1955, p. 265.
  15. Albert Einstein, “Elektrodynamik bewegter Korper,” Annalen der Physik, ser. 4, vol. 17, pp. 891-921.
  16. Thomas S. Kuhn, The Structure of Scientific Revolutions, 3rd edition, (Chicago: University of Chicago Press, 1996), pp. 72-75, 101-102.
  17. Albert Einstein, “Ist die Tragheit eines Korpers von seinem Energieinhalt abhängig?”, Annalen der Physik, ser. 4, vol. 18, pp. 639-641. The famous equation for the equivalence of mass and energy appears here in slightly different notation.
  18. On the early history of special relativity, see Pais, “Subtle is the Lord…”, pp. 138-162. The Planck quotation is found on p. 382.
  19. Ibid., p. 179.
  20. To return to the example of the railway car, the experimenter inside the car cannot tell whether the feeling of his own weight on the floor is due to an external force of gravity or to the upward acceleration of the entire railway car.
  21. A. Einstein, “Grundlagen der allgemeinen Relativitatstheorie,” Annalen der Physik, ser. 4, vol. 49 (1916), pp. 769-822. This was also published as a separate volume (Leipzig: Barth, 1916).
  22. Pais, ‘Subtle is the Lord…’, pp. 379-386.
  23. The best intellectual biography of Bohr is Abraham Pais, Niels Bohr’s Times: in Physics, Philosophy and Polity, (Oxford: Clarendon Press, 1991).
  24. Two historians of science had recently re-examined the Eddington data and have argued that the results are not as clear as Eddington claimed. Harry Collins and Trevor Pinch, “Two Experiments that ‘proved’ the theory of relativity,” The Golem: What You Should Know about Science, 2nd Ed., (Cambridge: Cambridge Universtiy Press, 1998), pp. 27-55, 155-180.
  25. For the politics behind Einstein’s Nobel Prize, A. Pais, “Subtle is the Lord…”, pp. 502-512.
  26. Morton Tavel, Contemporary Physics and the Limits of Knowledge, (New Brunswick: Rutgers University Press, 2002). Chapter 10 contains a clear description of quantum mechanics and of Einstein’s objections. Chapters 4 through 7 are also a good introduction to the special and general theories of relativity.
  27. Einstein to Born, December 4, 1926. Quoted in Pais, “Subtle is the Lord…”, p. 443.
  28. Ibid., pp. 440-449.
  29. Brian, op.cit., p. 425.
  30. Pais, Einstein Lived Here, pp. 142-145.
  31. Scientific American Supplement, Nov. 11, 1911. Quoted in Daniel C. Schlenoff, “A Century of Einstein,” Scientific American, Vol. 291, No. 3 (Sept., 2004), p. 103.
  32. Pais, Einstein Lived Here, p. 138.
  33. Ibid., p. 252.
  34. London Times, Nov. 7, 1919. The full article is reproduced in Pais, “Subtle is the Lord…”, p. 307.
  35. A. Einstein, “What is the Theory of Relativity?” The London Times, November 28, 1919. Reprinted in Albert Einstein, Ideas and Opinions, ed. By Carl Selig, (New York: Dell, 1954), pp. 222 - 227.
  36. New York Times, Nov. 10, 1919. The New York Times reporter in London, who usually covered golf, did not bother to attend the meeting, but based his story on the London Times account and a telephone interview with Eddington. Brian, op.cit., pp. 100-101.
  37. Quoted in Pais, Einstein Lived Here, p. 147.
  38. Ibid., p. 146.
  39. New York Times, April 3, 1921, p. 1. The front page is reproduced in Ze’ev Rosenkranz, The Einstein Scrapbook, (Baltimore: Johns Hopkins University Press, 2002), p. 114.
  40. Ibid. At this time Einstein spoke English only with great difficulty (and he was never comfortable lecturing or discussing science in anything but German) and had to rely on others to translate for him. This was certainly part of the reason reporters found him difficult to understand.
  41. Quoted in Pais, Einstein Lived Here, p. 184.
  42. Ibid.
  43. Ibid., p. 148.
  44. This is mentioned in the first paragraph of his 1919 article in the London Times. Einstein, Ideas and Opinions, p. 222.
  45. Quoted in Pais, Einstein Lived Here, p. 149.
  46. Pais, Einstein Lived Here, p. 150.