Albert Einstein
Albert Einstein; drawing by David Levine

1.

Until the age of seven or eight, whenever the young Albert Einstein was asked a question, he would slowly formulate an answer, mutter it tentatively to himself, and finally repeat aloud his considered response. This laborious method of speaking gave the impression that he needed to say everything twice. His parents consulted a doctor, and the family housekeeper called the boy “stupid.” Decades later, Einstein’s sister Maja recorded this odd childhood habit and attributed it to her brother’s thoroughness in thinking. Yet the doubling of each sentence, once for himself and once for everyone else, may also have been an early sign of the deep inner world that Einstein inhabited. Brilliant, supremely self-confident, brutally honest, witty, stubborn—Einstein was above all else a loner.

In an essay he published in 1931, at the age of fifty-two, the physicist wrote:

My passionate sense of social justice and social responsibility has always contrasted oddly with my pronounced lack of need for direct contact with other human beings and human communities. I am truly a “lone traveler” and have never belonged to my country, my home, my friends, or even my immediate family, with my whole heart; in the face of all these ties, I have never lost a sense of distance and a need for solitude—

With this passage one cannot help recalling a close contemporary of Einstein’s, the poet Rainer Maria Rilke, who famously advised another young poet to “love your solitude and bear with sweet-sounding lamentation the suffering it causes you.” Einstein’s isolation was surely in part the artist’s compulsion to create alone. And the great physicist was indeed an artist in his devotion to simplicity and mathematical beauty. But his distance went far beyond any aesthetic concerns. Throughout his life, he maintained a strong awareness that he did not fit in, intellectually, socially, spiritually. Einstein had a profound sense of otherness, even alienation.

Numerous anecdotes from childhood suggest that these feelings were partly a consequence of innate temperament. But they were also strongly accentuated by his harsh and authoritarian early teachers, the German military service that caused him to renounce his citizenship at the age of sixteen, his parents’ contempt for his sweetheart and first-wife-to-be, Mileva Mariå«c, his inability to secure university employment after college, and finally his growing identification with the plight of his fellow Jews, whom he referred to as his “tribal companions.”

The 1931 passage on solitude, together with many of Einstein’s other public essays, have long been available in two of his books, Mein Weltbild (1934) and Ideas and Opinions (1954). (The latter is an extraordinary compendium of Einstein’s thoughts on philosophy, religion, education, politics, and the methods of science.) These writings, from later life and after his rise to worldwide fame, conform more or less to the popular image of Einstein as a wise, grandfatherly figure. When hints of the lone trav-eler appear here and there, they are couched in rather abstract language. A more private and gritty view of Einstein is now emerging in several new biographies, stimulated by the recent availability of the younger Einstein’s personal letters and perhaps also by the recurrent fascination with the frailties of our heroes. Beyond the usual revealed intimacies and imperfections is clear evidence that Einstein’s sense of estrangement began at a young age.

Of the new biographies, the fullest by far is Albrecht Fölsing’s huge Albert Einstein, originally published in German in 1993 and now appearing in English.1 Fölsing’s book, while needing clearer scientific explanations, provides a nuanced, nonjudgmental personal portrait, resting firmly on prodigious archival work. Albert Einstein allows us to see deeply into Einstein’s inner world, to hear his voice, both muttering and proclaiming, and thus offers an excellent companion to Abraham Pais’s fine intellectual biography Subtle Is the Lord.2

Prime among the new source material is a set of fifty-two love letters exchanged between Einstein and Mileva Mariå«c, extending from 1897 to 1902. These letters were released to scholars only in 1986, upon the death of Einstein’s older son, Hans Albert. They have been printed in full in the first volume of The Collected Papers of Albert Einstein, an immense project jointly sponsored by the Hebrew University of Jerusalem and Princeton University Press. (At his death in 1955, Einstein willed all of his papers and letters to Hebrew University; the Einstein Archive there now contains some 45,000 documents, and The Collected Papers is expected to run to over twenty-five volumes.)

Mileva Mariå«c, born in 1875 and three years older than Albert, came from a well-to-do Serbian family living in what was then Hungary. Darkly pretty, with a slight limp from childhood, she excelled in physics and mathematics. At the time that her surviving correspondence with Albert began, in October 1897, they were both in their second year of study at the Polytechnic in Zurich—Switzerland being the only German-speaking country that then allowed women into higher education. Mileva had just taken a temporary leave to attend lectures at the University of Heidelberg.

Advertisement

From these letters we learn in Einstein’s own words how he is tormented by his parents over Mariå«c. In 1900, he writes her: “My parents are very distressed about my love for you, Mama often cries bitterly & I am not given a single undisturbed moment here. My parents mourn for me almost as if I had died.” A hint of Mama’s objections is given in a letter Einstein wrote to Mariå«c a few months earlier, in which he quotes his mother as saying: “She is a book like you—but you ought to have a wife.” “When you’ll be 30, she’ll be an old hag.”

Other letters convey a strong sense that the young physicist was battling the world, not just his parents. In 1901, Mariå«c failed her degree examination at the Polytechnic and moved back with her parents while she was pregnant with Einstein’s illegitimate child. (The child, a daughter named Lieserl, was evidently given up soon after birth; no records have been found, and she has vanished from history.) Einstein, meanwhile, was living close to poverty and having no luck in obtaining a job. It is under these circumstances that he writes to her:

I decided the following about our future: I will look immediately for a position, no matter how humble…. The moment I have obtained such a position I’ll marry you and take you to me without writing anyone a single word before everything has been settled. And then nobody can cast a stone upon your dear head, and whoever dares to do anything against you, he’ll better watch out!

Then, in December 1901, just after Einstein has learned that he may be offered a clerkship in the government Patent Office in Bern, he writes: “We shall remain students (horribile dictu) as long as we live, and shall not give a damn about the world.” Two weeks later, he writes: “Apart from you, all the people look so alien to me as if they were separated from me by an invisible wall.”

The young Einstein was especially embittered by his failure to receive recognition from the academic establishment, many of whose eminences he considered self-satisfied men, far beneath him in scientific ability. After sending a letter to Paul Drude, a leading physicist and editor of the prestigious Annalen der Physik, about some errors in Drude’s work, the twenty-two-year-old Einstein received an unyielding reply from Drude, and wrote to Mariå«c in July 1901:

I have just come home from Lenzburg & found this letter from Drude, which is such an irrefutable evidence of its writer’s wretchedness that no comment by me is necessary. From now on I’ll not turn any longer to this kind of person but will rather attack them mercilessly via journals, as they deserve. It is no wonder that little by little one becomes a misanthrope.

In November 1901, Einstein submitted a doctoral thesis to Professor Alfred Kleiner at the University of Zurich, criticizing some of the work of the great Ludwig Boltzmann, a colleague of Kleiner. In December of that year, Einstein wrote to Mariå«c:

Since that bore Kleiner hasn’t answered yet, I am going to drop in on him on Thursday…. To think of all the obstacles that these old philistines put in the way of a person who is not of their ilk, it’s really ghastly! This sort instinctively considers every intelligent young person as a danger to his frail dignity, this is how it seems to me by now. But if he has the gall to reject my doctoral thesis, then I’ll publish his rejection in cold print together with the thesis & he will have made a fool of himself.

Albert Einstein was born on March 14, 1879, in Ulm, Germany. His mother, Pauline Koch, was musical and helped get young Albert started in playing the violin. His father, Hermann, was in the electrical business, first with his younger brother Jakob and later on his own, initially investing his own money and then borrowing family money. Hermann suffered one business failure after another. In the hopes of achieving success, the family moved several times, first to Munich in 1880, then to Milan in 1894, to Pavia in 1895, and back to Milan in 1896. Einstein entered primary school in Munich, at the age of six. As his sister Maja recalled in her unpublished 1924 biography of her brother, Einstein’s teacher taught her students the multiplication tables with the help of whacks on the hand, all to better prepare them early to be good German citizens. Einstein was the only Jew among seventy classmates. One day his teacher of religious studies brought a nail to class and told his students that with such nails Christ had been nailed to the cross by the Jews.

Advertisement

At home, young Albert was becoming more and more of an introvert. Instead of playing with other boys, he liked to work out puzzles, to create complicated constructions out of building blocks, and, with great patience and determination, to build houses of cards many stories high. In the fall of 1888, the boy entered secondary school, the Luitpold Gymnasium. When his family moved to Milan in the summer of 1894, the fifteen-year-old Einstein stayed behind in Munich to finish his studies, boarding with a family. In December, after an argument with one of his teachers at the gymnasium, Einstein abruptly dropped out of school and moved back to Milan with his family. He announced to his parents that he would never return to Munich and would prepare himself to enter college at the Zurich Polytechnic. Then, and for the next few years, he voraciously read physics and mathematics on his own.

After graduation from the Polytechnic in 1900, Einstein applied for the position of assistant to most of the leading physics professors in Europe. He was turned down by all of them, possibly because of cool letters from his professor at the Polytechnic, Heinrich Friedrich Weber, whom Einstein admired but did not lavish with the customary subservience. (Mariå«c quietly offered her explanation in a letter to a friend: “You know that my sweetheart has a sharp tongue and moreover he’s a Jew.”) Einstein barely supported himself by tutoring for two years before he received his job at the Patent Office in Bern, in 1902.

On January 6, 1903, Einstein and Mariå«c were married in the stark registry office of Bern—no wedding guests, no members of either family. Serving as witnesses were two friends of the groom, Conrad Habicht and Maurice Solovine, fellow members of the self-appointed “Olympia Academy.” For several years, the three friends had been meeting a few nights a week at one of their houses for sausage, cheese, and discussion of Hume, Kant, Dickens, Poincaré. In November, the newly married couple moved to a rented apartment at 49 Kramgasse, in the old city of Bern. The apartment, on the third floor and reached by a narrow staircase, had only two rooms. However, one of the rooms had large windows looking down on the fine cobblestone street and quaint shops below.

It was here, in the single year of 1905 and working entirely on his own, that the twenty-six-year-old physicist produced four papers which changed the course of physics. Each of the treatises was sufficient to earn him lifetime recognition; one brought him the 1921 Nobel Prize. Two of them presented a mathematical analysis of Brownian motion, the irregular movement of minute particles suspended in a fluid, and provided definitive evidence for the existence of molecules. Another of his papers, which he described to a friend as “very revolutionary,” suggested that certain recent experiments could be understood if light flowed in discrete packets of energy, like water droplets, rather than in a continuous stream. The fourth, on Special Relativity, the greatest achievement of all, proposed a new understanding of time and space.

Only months after these papers appeared in the Annalen der Physik, the unknown man at the Patent Office in Bern was receiving admiring letters from the likes of Phillip Lenard and Max Planck, respectively the leading experimental physicist and the leading theoretical physicist in the German-speaking world. News of the clerk’s original thoughts traveled with less velocity in Bern. In April 1906, Einstein’s salary at the Patent Office was increased from 3900 to 4500 francs. The physicist attempted but failed to get a better job at the Bern Post and Telegraph Directorate.

Within a few years, however, he received a succession of professorships. In 1909 Einstein was appointed professor at the University of Zurich (where his work on Brownian motion had been accepted for a doctorate four years earlier) and he received his first honorary degree from the University of Geneva. In 1911, Emperor Francis Joseph appointed Einstein a full professor at the German University in Prague; the next year he moved back to a professorship at the Zurich Polytechnic, now called ETH; in 1913 Kaiser Wilhelm II confirmed his appointment to the Prussian Academy of Sciences in Berlin. From then on, Einstein did little formal teaching, which he disliked; in contrast to almost all other academic physicists, great and not so great, he supervised only a single doctoral student in his life. Einstein worked mainly in Berlin until 1933, when he took up his final post as a professor (without duties) at the new Institute for Advanced Study in Princeton.

Since his marriage in 1903, Einstein had been a remote and insensitive husband and father, refusing to take on family responsibilities that might interfere with his work. Months after moving to Berlin, he separated from Mariå«c, who returned to Zurich with their two sons, Hans Albert and Eduard. The divorce settlement in 1919 must be unique in the history of domestic conflicts. In addition to giving Mariå«c custody of the children, Einstein promised her his putative Nobel Prize money, which he and all other physicists on the planet were certain would be awarded. (Such consensus on achievement may be possible only in science.)

The Nobel arrived in 1922. At that time, a journalist gave the following wonderful description of Einstein:

Einstein is tall (about 1.70 m), with broad shoulders and a scarcely bent back. His head—the head in which the science of the world was newly created—instantly attracts lasting attention…. A little mustache, dark and very short, adorns a sensuous mouth, very red, rather big, with its corner betraying a permanent slight smile. But the strongest impression is that of stunning youthfulness, very romantic and at certain moments irresistibly reminiscent of the young Bee-thoven who, already marked by life, had once been handsome. And then suddenly laughter erupts and one is faced with a student.

Einstein’s name was now a household word. In fact, he had rocketed to international fame a few years earlier, in 1919, when observations during a solar eclipse confirmed that light rays from stars were deflected by the gravity of the sun to just the extent he had predicted with his new theory of gravity, General Relativity. General Relativity was decades ahead of other physicists. It envisioned gravity as a warping of space, the curved orbits of planets as longitudes on a twisting cosmic terrain. Headlines appeared throughout the world. The London Times of November 8, 1919, for example, announced: “The Revolution in Science/Einstein Versus Newton.” The world was exhausted by World War I, eager for some sign of humankind’s nobility, and suddenly here was a scientific genius, seemingly interested only in purely intellectual pursuits. Almost overnight, Einstein became the most honored scientist in history. The new Isaac Newton was invited to address congresses and leaders of state throughout the world; he was given banquets and accolades wherever he went; he was interviewed ceaselessly by the press; great intellectuals and artists, such as Rabindranath Tagore and Sigmund Freud, visited or corresponded with him.

The world was soon to discover that Professor Einstein was not a purely cerebral academic but a vocal intellectual, worthy of comparison to Voltaire as well as to Newton. Beginning in the early 1920s, Einstein drew the hatred of many in his fatherland by his many commitments to the political left. He supported the League for Human Rights and, in 1921, helped found the Society of Friends of the New Russia. In 1928, his early pacifism turned militant when he publicly criticized the mostly impotent League of Nations for codifying the rules of warfare: “War simply is no game and cannot be conducted according to the rules of a game. War must be opposed as such, and this can be done most effectively by the masses through an organized wholesale refusal of military service already in peacetime.” Two years later, from aboard a ship sailing into American waters, he gave his famous “Two percent speech”: “If even two percent of those called up declare that they will not serve, and simultaneously demand that all international conflicts be settled in a peaceful manner, governments would be powerless.”

Many Germans of the late Weimar period, particularly those preparing to launch a new war, smoldered with anger at these flagrantly unpatriotic statements from the most celebrated German in the world. Yet their hatred could not break into the open, for they needed Einstein more than he needed them. Some of the leading scientists were naturally anxious that he remain in the country. The great physicist Max Planck constantly pleaded with Einstein not to resign his seat in the Prussian Academy of Sciences and many others in high German society continued to praise him as an exponent of German culture. To all of which Einstein responded with characteris-tic bite: “A funny lot, these Germans. To them I am a stinking flower, and yet they keep putting me into their buttonhole.”

Finally, in 1933, Einstein could no longer allow himself to be stuffed into the German buttonhole and left forever for the United States. He was being denounced in Germany not just for his pacifism but also for his religion. Einstein had never made much of his Jewishness until he went to Berlin and first “understood the Jewish community of destiny,” united by a common history of persecution. For Einstein, that common persecution connected with his own life experiences, religious and secular. To a group of Jewish students he later said: “We should be clearly aware of our otherness and draw our conclusions from it. There is no point in trying to convince the others of our emotional and intellectual equality by way of deduction, since the root of their behavior is not located in the cerebrum. Instead we should socially emancipate ourselves and essentially satisfy our social requirements ourselves.”

In 1935, Einstein and his second wife, Elsa Löwenthal, moved to an inconspicuous house at 112 Mercer Street in Princeton, a New England clapboard with a small garden in front and a long narrow garden in back. Einstein took as his study a room on the second floor, with a large picture window that looked across a wooded park to the university’s gothic buildings in the distance. It was here that Einstein spent most of his time, and was the happiest, in the last two decades of his life.

2.

At the turn of the century, when the young Einstein began crafting the theory of Special Relativity, physicists knew of two fundamental forces of nature: the gravitational force and the electromagnetic force. (Electrical and magnetic forces were considered parts of the same force because they could generate each other, as a wire twirling near a magnet generates an electrical current.) Einstein did not initially set out to propose new concepts for time and space but rather to explain a particular problem in the theory of electromagnetism.

The problem was this: during the last decades of the nineteenth century, physicists had shown that light consisted of undulating electrical and magnetic energy traveling through space. All known undulations of energy, such as sound waves, required a material substance to traverse. For example, when sound travels across a room, it is the air molecules in the room, bumping successively into one another like a row of falling dominoes, that constitute the traveling “wave” of sound. Sound can travel through a gas, a liquid, or a solid, but it cannot travel through a vacuum. It needs the dominoes. By analogy, scientists assumed that the propagation of light required an underlying material substance. This substance was called the ether. Furthermore, the ether, a gossamer and practically weightless substratum, was required to fill up every nook of the universe, since the received glimmer of distant stars clearly showed that light could travel through the vast reaches of space as well as across a living room.

Now, it is both intuitively obvious and also true that sound will travel more quickly across a room if it rides not on static air but on a wind blowing through the room. By analogy, physicists assumed that the speed of light should depend on the motion of the ether going past the observer or the observer going past the ether. (In physics, an “observer” is anyone equipped to make measurements.) The great quandary was that all attempts to measure the motion of an observer through the ether had proved unsuccessful. Light rays seemed to travel at the same speed regardless of which way the assumed ether “wind” was blowing. Electrical and magnetic experiments performed in a train station gave results identical to the same experiments performed in a train traveling by the station at 60 miles per hour.

The ether’s physical reality had never been confirmed. Einstein postulated that it did not exist, that it was a superfluous construction. Without an ether, it could not be said that an observer was moving or not with respect to the ether. In fact, it could not be said that an individual observer was moving at all, in absolute terms. By postulate, Einstein had eliminated the cosmic frame of reference against which all individual motions could be measured. Only the relative motion of two observers, or any two things, had meaning. Hence the name “relativity.” Correspondingly, all observers could be considered as equivalent. All observers, as long as they were not accelerating, would produce identical measurements for all electrical and magnetic phenomena. In particular, a man sitting on a couch would measure exactly the same speed for a passing light ray as a man running by him in the direction of the ray.

To permit himself this last seemingly preposterous proposition, Einstein first recognized that our notions of time and distance, both of which enter into the measurement of speed, had not been carefully analyzed before. Anchored in the unconscious, for example, was a belief in the absolute character of time. A second for me is a second for you. As a practical matter, the possible discrepancies between clocks moving past each other at the slow speeds of earthly life would be much too small to have been noticed. Philosophically, Einstein disagreed with Kant, whom he had read as a youth. He opposed Kant’s view that particular concepts of time and space are necessary premises for thought, innate in the human mind. For Einstein, all concepts, even physical ones, are “free inventions of the human intellect,” which can be modified if they do not prove useful in comparison to experience. The meaning of space and time (for physics) is determined, he wrote, by considering the spatial separation and time elapsed between two events, like two successive chimes of a grandfather clock. When observers in relative motion measure these same two events, their measuring sticks and clocks must be calibrated and synchronized in a reproducible way.

Starting with these working definitions of time and space, and the postulate that there was no ether, Einstein calculated quantitatively how the ticking rates of clocks and the lengths of measuring sticks in motion with respect to each other must differ so that both sets of instruments measure the same speed for a passing ray of light. The predicted differences are indeed tiny for everyday relative motions—one second on a supersonic jet takes 1.0000000000014 seconds as measured on the ground—but the differences have indeed been detected and agree with the theory. For relative speeds closer to the speed of light, 186,000 miles per second, such as occur in high-energy particle accelerators, the effects are much larger. A second for me is not a second for you. However, the deep significance of Special Relativity extends far beyond its quantitative results. More than any philosopher or scientist before him, Einstein demonstrated that our intuition about the physical world based on sensory experience can be fundamentally in error.

How much did Einstein’s temperament, especially his sense of being an outsider, affect his science? Here, of course, one can only point to certain suggestive associations rather than prove causal relationships.

First of all, Einstein used a rare, deductive approach to science. In this he differed from most other scientists of the time, including other theoretical physicists. In the more usual inductive approach, the scientist begins with a number of observations about nature, tries to find a pattern, generalizes the pattern into a “law” or organizing principle, and then tests this law against future experiments. For example, the German astronomer Johannes Kepler examined the data on planets, analyzing the numbers in many different ways, before discovering a striking relationship between a planet’s distance from the sun and the time it takes to complete an orbit. This brief account of the methodological process is certainly oversimplified, and the work is often carried out by an entire community of scientists rather than by a single person. Essentially, however, experiments, data, and observations form both the starting point and the center in inductive science. In deductive science, on the other hand, the scientist begins by postulating certain things to be true, with only secondary guidance from outside experiments, and then deduces the consequences of those postulates. Finally, the consequences are tested experimentally. If the tests fail, the postulates must be changed.

In his Autobiographical Notes (1946) Einstein explicitly credits the philosopher David Hume for having taught him that the truths of nature cannot be arrived at by experience with the world. Rather, one must start with the “free inventions” of the mind. In the case of Special Relativity, for example, Einstein began with the postulate that the ether did not exist and that absolute motion did not exist; he then derived the consequences of that idea. Other theoretical physicists of the time, most notably the Dutch Hendrik Antoon Lorentz, proposed detailed theories for how bodies moving through the ether would be electrically compressed in exactly the right amount, depending on their speed, in order to make the experimental results come out as they did. It seems possible that the depth of Einstein’s inner world made him especially attentive to Hume’s teachings and naturally inclined to explore nature from within his own mind.

Two aspects of Einstein’s later scientific career are suggestive of his stubborn self-confidence and willingness to strike out completely on his own: his philosophical refusal to accept the new quantum physics, and his continued work on his own Unified Theory (combining gravity and electromagnetism), despite one failure after another.

A fundamental tenet of quantum theory, developed in the 1920s, is that the position and velocity of an individual particle cannot be completely specified, even in principle. As a result, one cannot predict with certainty the future position and velocity of a particle; such predictions can be done only in terms of probabilities, which apply to the average behavior of a large number of particles. Einstein stridently opposed the indeterminacy inherent in quantum theory. In his view, everything, even the behavior of a single electron, should be deterministic and calculable. In a letter to his fellow physicist Max Born, Einstein wrote: “The idea that an electron exposed to a ray by its own free decision chooses the moment and the direction in which it wants to eject is intolerable to me. If that is so, I’d rather be a cobbler or a clerk in a gambling casino than a physicist.”

While the current of physics swerved in the direction of the powerful new quantum physics, which was able to explain the behavior of atoms, Einstein became obsessed with his nonquantum Unified Theory, which he doggedly pursued in increasing isolation for the rest of his life. With each new attempt to perfect that theory, the physicist was sure that he was closer to victory. (“The latest results are so beautiful that I have every confidence….”) The discovery of antimatter in the 1930s and of new subatomic particles, hinting at new fundamental forces, all passed by with little comment from Einstein. Even when the great Danish physicist Niels Bohr, with whom he had once had lively discussions, visited the Institute for Advanced Study in 1939, Einstein remained cloistered in his paper-strewn room at the Institute.

Einstein’s brittleness here cannot be attributed simply to age. Born and Bohr, for example, were only a few years younger than Einstein, yet both immediately embraced the new quantum theory and became leading practitioners of it. In a letter to his friend and former physician Otto Juliusburger in 1937, Einstein wrote: “I am not really becoming part of the human world here [in Princeton], for that I was too old when I arrived, and in point of fact it was no different in Berlin or in Switzerland. One is born a loner.” Einstein’s independence and stubbornness, which had served him so well in the first part of his scientific career, became a curse in the last.

Most striking about Fölsing’s rich biography is its portrait of a man of contradictions. Einstein scoffed at material concerns, yet he consistently demanded exorbitant salaries and lecture fees once he became famous. He opposed nationalism of all kinds, including Jewish nationalism, but he supported the birth of Israel and later the founding of the Hebrew University. He was an adamant pacifist who scorned the German military mentality. Yet he remained friends with Fritz Haber, who inaugurated chemical warfare and the use of poison gas in 1915 (the year Einstein published his paper on General Relativity), and in fact for some time maintained his office in Haber’s Kaiser Wilhelm Institute. To his friend H. Zangger, Einstein wrote in 1917: “Our entire much-praised technological progress, and civilization generally, could be compared to an axe in the hand of a pathological criminal.” Yet during this same period Einstein worked on the improvement of aircraft wings and collaborated with the rich young inventor Hermann Anschütz-Kaempfe on the first gyrocompass, used on German U-boats during the war. (By the 1930s, almost every navy in the world, except the British and American, steered their courses with gyrocompasses based in part on a patent owned by Einstein.)

Einstein the humanitarian lent his name to social causes and, once the great wave of European Jewish emigration began in the mid-1930s, he personally supported and financed the families of German friends in their relocations in America. After the German annexation of Austria in 1938, he tried to start a relief organization supported by churches, universities, and the Red Cross. Yet in other circumstances the same man seemed an elitist. To his cousin Elsa, whom he was to marry six years later, he wrote in 1913: “Traveling people, both of us, destined to high-wire dancing from among the swarm of Philistines…,” and “I have firmly resolved to…walk only in really agreeable company, in other words rarely.” In 1931, he wrote of the death penalty: “I would have no objection to the killing of worthless or even harmful individuals; I am against it only because I do not trust people, i.e., the courts. I appreciate more the quality than the quantity of human life.” In the publication Forum and Century the same year: “The really valuable thing in the pageant of human life seems to me…the creative, sentient individual…while the herd as such remains dull in thought and dull in feeling.”

Einstein was both a man of high principles and an opportunist, a loner and an activist, a liberal and an elitist, a great theoretician and a practical examiner of patents. By the time he was forty, he was well aware of his high place in history. Yet he showed himself capable of humility and kindness. “A hundred times every day I remind myself that my inner and outer life are based on the labors of other men, living and dead,” he once wrote. In Einstein’s last decade of life, his ophthalmologist in Princeton checked his eyes annually and invariably told the old man that his glasses would be much improved with a new prescription. And Einstein would invariably reply with a smile: “A friend in New York sends me these simple magnifying glasses as a gift each year, and if they do no real harm, Henry, I prefer not to change them. I don’t want to hurt his feelings.”

This Issue

April 10, 1997