Leo Szilard
Leo Szilard; drawing by David Levine

I suppose that a figure could be put to the number of books that tell the story of the unfolding of our nuclear age. There must be hundreds about the development of atomic weapons, and no doubt as many more will be published as further chapters are added to the story. This is not to say that we can all follow what we read about the way the anatomy of the atom was laid bare. What is important is that all of us know that the physicists were not joking when they said that if an atom were split—whatever the term meant—the energy that had previously been bound within the subatomic particles of which it consists could be released, as Hiroshima and Nagasaki were to bear witness, as a fabulously destructive force. We can be sure that they are not exaggerating now when they tell us that it is possible to pack sufficient destructive power into a single hydrogen bomb—say, one of the many that could get through President Reagan’s SDI defensive screen—to crush and burn to a frazzle all the inhabitants of Washington, D.C., even all those of the whole of Manhattan. Winston Churchill was on sound ground when he said that the bomb would carry “mankind outside the scope of human control.”1

Obviously that was not what the physicists and mathematicians of the early years of our century were out to achieve. They were striving to penetrate the inner structure of the matter that our senses perceive. Their purpose was to add to the fund of basic knowledge, not to make us either materially richer or politically more powerful. The handful of scholars who had the brains, the imagination, and the courage to try to follow the pioneers went to Berlin, Copenhagen, Cambridge, and to other strongholds of the new physics, moving as peripatetic students from one center to another, until the moment came during the early Thirties when they were stopped by the barrier of fascism.

Among those who joined the freemasonry of the new physics were Isidor Rabi, Leo Szilard, and Edward Teller. All were born in central Europe, all were to play a part in the story of “the bomb,” and all were to have a profound impact on the history of their adopted country, the United States, and indeed of the world. Their professional paths were to converge, for a time coalesce, until the day when one, Teller, struck out in a direction diametrically opposed to that of the other two. The close appearance of the three books with which this review deals is a challenge for us to try to understand the curious relations these men had with each other.

Isidor Rabi, the most distinguished of the three, died this January. He was born in 1898, in a small town in what is now Poland, into a poor and strictly Orthodox Jewish family that emigrated to the United States before he was a year old. What happened next is told in John Rigden’s fascinating book, based as it is on wide reading, on interviews with more than a score of scientists who worked with Rabi, and, most important, on extensive talks with Rabi and his wife.

Rabi’s formal education started in a small Hebrew school in New York, where, without ever abandoning a sense of religious belief, of “the mystery and the philosophy” of the Creation, he started to question Jewish orthodoxy. A local public library had introduced him to a wider world. The Copernican solar system, about which he read in a little book on astronomy, came as a revelation that eliminated his need to believe in a God whose job it was to make the sun rise and set every day. At his next elementary school, he read about socialism and Marxism, which he soon realized did not provide the integrated view of society that he was seeking—for by the age of thirteen he already felt that there had to be some conceptual scheme into which could be fitted the disparate events of the past and present. He built his own radio set, and startled his elders by taking as the theme of the speech that by convention he had to deliver at his bar mitzvah ceremony “How the electric light works.”

The high school where he spent the next four years was a craft school, chosen as a way of escaping from his otherwise narrow environment. He left it “streetwise and self-educated”—streetwise because he had to learn to avoid the gangs that set upon small Jewish boys on their way to and from school; self-educated, because the school left him all the time in the world for his own reading. A scholarship took him to Cornell, which he entered as a student of electrical engineering. He quickly switched to chemistry—and continued the process of self-education through avid reading of whatever caught his fancy. In those days the atmosphere of Cornell was so anti-Semitic that during Rabi’s four undergraduate years he did not get to know a single faculty member. Nor, after he graduated in chemistry in 1919, were there any jobs available either for him or his Jewish classmates. He spent the next three years under the parental roof in New York, moving from one odd job to another, continuing with his reading, and enjoying the close company of three friends who had also graduated with distinction, but who were now in the same boat as he was.

Advertisement

This aimless period ended when he decided to return to Cornell as a graduate student, with his interest now turning to physics. But having failed to secure a fellowship with which to pay his way, he returned to New York, and entered Columbia University. A part-time tutorship at City College provided him with a faint measure of financial security. The subject of his thesis for the Ph.D., which he was awarded in 1926, was the magnetic susceptibilities of crystals, a problem that turned his attention to quantum mechanics. Only the new physics could explain why different atoms have different chemical properties.

A fellowship that he was awarded in 1927 then allowed him to embark on a tour of the European centers where the new subject was flourishing. In Munich, he ran into other American scholars who had set out on the same pilgrimage. He met young European theoretical physicists, among them Hans Bethe and Rudolf Peierls. He moved to Niels Bohr in Copenhagen, and then to Wolfgang Pauli in Hamburg. Here he stayed, doing brilliant work in a stimulating intellectual environment. For a short time he studied with Werner Heisenberg in Leipzig, where he met both Robert Oppenheimer and Edward Teller. His final port of call was Zurich, to which Pauli had moved from Hamburg, and where he ran into still more of the pioneers of the new physics, including Szilard and two other Hungarian scholars, Eugene Wigner and John Von Neumann. After some three months there, he returned in 1929 to Columbia, where he had been appointed to a lectureship.

During the next decade, Rabi’s career followed a much more even and far more successful course than did that of either Szilard or Teller. With his reputation as a theoretical physicist already made, he had been appointed to Columbia on the understanding that he would introduce quantum mechanics into the physics curriculum. But Rabi the theorist soon became Rabi the experimentalist. He explored the properties of atomic nuclei by ingenious developments of the molecular-beam technique that he had learned in Stern’s laboratory in Hamburg. The researches which he and his postgraduate students completed during the few years that preceded the Second World War were so outstanding that they not only established him as one of the great masters of his subject, but also won him, in 1944, the Nobel Prize for physics. By this time, he was deep in wartime research, working both on the development of radar at MIT and the atomic bomb at Los Alamos.

Rabi’s achievements were to be later memorialized in a schematic picture of a tree (reproduced by Rigden) which, as it grows, throws out powerful branches from which hang an ever-increasing number of basic scientific discoveries. The Rabi tree takes the story of his impact on the development of physics only to the end of the Fifties. By then it already recorded the names of twenty other Nobel laureates, including a few who had been his pupils. In a tribute to Rabi on his eightieth birthday, it was justly said that he had not only greatly enriched the cultural and intellectual life of Columbia, but had helped to establish the university as a “foremost center of scientific teaching and research.”2

Leo Szilard’s life, as Barton J. Bernstein reveals in his lively introduction to Toward a Livable World, a collection of papers, letters, and documents, was vastly different from Rabi’s. Szilard was born in Budapest in 1898—the same year as Rabi—the son of a prosperous Jewish civil engineer and builder, and was educated at Budapest’s premier Gymnasium, the Minta, where he is described as having been both bright and popular, and from which he graduated in 1916, winning the Hungarian national prize for mathematics. The First World War was on, and no sooner had he entered the university as an engineering student than he was drafted into the Austro-Hungarian army, from which however he was soon demobilized. When Béla Kun’s short-lived Hungarian Communist revolution broke out in 1919, Szilard decided that it was time to leave Budapest and to move to the University of Berlin.

Advertisement

He had sensed that worse was to come. In those days Jews made up only 5 percent of Hungary’s population of some 20 million, yet they filled between 50 and 60 percent of the professions and, almost more important, controlled 80 percent of the country’s financial institutions. They were powerful in the press and, although only a few were farmers, had come to own 40 percent of the country’s arable land. There were Jewish radicals as well as court Jews—financiers to a Magyar landowning nobility that had managed to keep a third of the country’s population in abject and illiterate poverty. The country was ripe for revolution. As a small but highly successful minority, exercising far more political power than was commensurate with their numbers, the Hungarian Jews were living on the slopes of an anti-Semitic volcano.3

In Berlin, Szilard abandoned the idea of a career in engineering and switched to physics. Never backward, he introduced himself to Einstein, and became a student of Max von Laue, the founder of X-ray crystallography. A highly original and quickly completed piece of research in the field of thermodynamics gained him his Ph.D. He was then appointed a lecturer in the University of Berlin—an appointment that he held until 1933 when, with the advent of Hitler, he left for London.

Szilard was an engaging extrovert, an excellent and humorous conversationalist, keenly interested in what was going on around him. Bubbling with ideas, and interested as much in politics as in science, he started to dream up ways of saving the world, one of his notions, so Bernstein tells us, being that of establishing an elitist and altruistic “Bund” of intellectuals which would take over the management of society when parliamentary bureaucracy collapsed. It sounds a little like the imaginary House of Solomon, which the great Francis Bacon, one of the founders of modern science, had sketched out in the early seventeenth century. The wise men who peopled Bacon’s house would themselves decide what information could be usefully and safely passed on to those who were overtly in executive charge of what went on in society. Szilard seems to have added a step to Bacon’s idea. On a more practical note, he is said to have kept a couple of suitcases ready for a quick getaway. He had long sensed that the Weimar Republic could not last.

When he reached London, he got in touch with a group of prominent British scholars who had formed an Academic Assistance Council to help university refugees who were fleeing from Hitler’s Germany. Szilard’s work was already known, and although he did not immediately seek an academic post, he continued to follow all that was going on in his subject. The neutron had just been discovered, and artificial radioactivity demonstrated. He conceived the idea that nuclear energy could be harnessed, and that it should therefore be possible to design a nuclear bomb. If neutrons could be “fired” into an appropriate element, they would split the atoms they struck, and in so doing release more neutrons to split other atoms. In 1934 Szilard took out two patents, which he assigned to the British Admiralty, describing the “laws” that would govern a “chain reaction” in a “critical mass” of a suitable substance. He was obsessed with the fear that the Germans might get there first.

Helped by an obliging biophysicist in a London teaching hospital, he then carried out a few preliminary experiments, the results of which convinced him that he was on the right track. The techniques he devised were essentially those that were later independently perfected by Enrico Fermi and Otto Frisch.4 In September 1935 he reported his results at a scientific meeting and attracted the attention of Professor F.A. Lindemann, the head of the Oxford physics department. Lindemann, who was to become better known as Lord Cherwell, Winston Churchill’s personal scientific adviser, offered him one of a few research fellowships that he had been instrumental in getting the chemical firm ICI to establish.

His department in Oxford was nowhere near as famous as was Rutherford’s in Cambridge but, helped by Szilard and two or three other refugee scientists, Lindemann soon assembled a research team that worked on particle physics. Szilard, however, was lonely. As he wrote, “being at Oxford but not of Oxford leaves you out of things.” In those days there were many Oxfords within Oxford—maybe there still are. I was a member of the same High Table as Lindemann but, to the best of my memory, he never once brought Szilard to dine with us. It was only years afterward that I met him.

In 1938 Szilard visited the United States, intending to stay only a short time. It was to become his permanent base. Toward a Livable World is essentially a record of his single-minded postwar crusade for world peace, and an end to the nuclear arms race. It consists of a selection of his vast correspondence with politicians and scientists, including a fascinating exchange with Khrushchev; drafts of memoranda addressed to the public, and records of interviews and of two important debates with Teller.

Edward Teller’s book, Better a Shield than a Sword, is an autobiographical compilation of a number of the author’s recent articles. In order to obtain biographical information, it is necessary to dig into Teller’s 1962 autobiography5 and into the hagiographical biography published in 1976 by Blumberg and Owens (who obtained most of their information from Teller himself).6

Like Leo Szilard, Teller was born in Budapest (in 1908) into a middle-class Jewish family. His father was an attorney, and Edward was spoiled mercilessly by his parents. “Please don’t talk to me,” the little boy would demand at the table. He was, he would say, working out a sum in his head. And the family would comply. After primary school, he was sent to the same Gymnasium that Szilard had attended ten years before. But from what one reads, it would appear that whereas Szilard was jolly, modest, and popular, Teller was opinionated, unpopular, and bullied by the other boys; he was too ready to show that he knew better than his teachers. His years at school were also caught up in the turmoil that followed the breakup of the Austro-Hungarian Empire. After Béla Kun’s shortlived Communist revolution came Admiral Horthy’s fascist regime, with its overt anti-Semitism. For a time soldiers were billeted in the Teller home. There were riots, and Jews lived in fear. Szilard had got out in time. The Teller family stayed put.

It was not until 1926, when he was seventeen, that Teller left Budapest to continue his studies of physics in Germany, first in the Technical Institute of Karlsruhe, from which he went, in turn, to Munich—where he lost a foot in an accident—and then to Leipzig, where he studied with Heisenberg, the theoretical physicist who was to remain at work in Hitler’s Germany. Having gained his Ph.D., Teller moved to Göttingen. Here he spent two years. But by then Hitler’s storm troopers were being mobilized. He left for Niels Bohr’s institute in Copenhagen, and while there he was awarded a Rockefeller Fellowship. He worked for a short time in London before migrating in 1935 to join the small community of nuclear physicists that was coming together in the United States. He was appointed to a professorship in physics at George Washington University, joining George Gamow, another émigré physicist.

During the four years that preceded the outbreak of World War II, Teller lectured to undergraduates, carried out research on thermonuclear energy, and helped to organize annual colloquia on theoretical physics to which he and Gamow invited a small and intimate group of physicists from other universities—among them some of their fellow émigrés. Teller’s scholarly arrangements were, however, soon to be shattered by his older co-patriot, Leo Szilard.

Szilard had arranged that the research fellowship that Lindemann had provided him could be spent partly at Oxford and partly in the US. The Hitler-Chamberlain Munich agreement coincided with his 1938 visit to New York, and convinced him, even more than he already was, that war with Germany was inevitable. Galvanized by the news that nuclear fission had been successfully demonstrated in uranium—first in Berlin and Copenhagen, and then in Paris—he also became more certain than ever of the validity of his theory that a “chain reaction” could be induced in a “critical mass” of uranium. Since there was a better chance of putting his ideas to the test in the United States than at Oxford, and having conferred urgently with Rabi and with the Italian physicist Enrico Fermi, who had joined Rabi at Columbia in early 1939, he and Fermi started some experiments. Szilard recruited Teller to assist them. One result was the famous letter that Einstein sent to President Roosevelt in the summer of 1939, warning the President that it might be possible to develop an atom bomb of unimaginable destructive force. The letter had been drafted by Szilard, who had then called upon Teller and Wigner to drive him to see Einstein.

Action followed, slowly at first, but quickly accelerating when, in the summer of 1940, the President established a National Defense Research Committee to expedite and coordinate all military research that was held to be relevant to American security. A uranium committee was set up to encourage research into nuclear fission. After Pearl Harbor in December 1941, Chicago was for a time made the center for the studies that the Uranium Committee supported. Then, early in 1943, Robert Oppenheimer was put in charge of a group of physicists to formulate precise proposals for the work that would have to be done to develop an atomic bomb. Teller, who had moved with Szilard and Fermi from New York to Chicago, was one of the physicists whom Oppenheimer invited to the discussions. Later that year a single administration, code-named The Manhattan Project, was made responsible for all research and production relative to the bomb, and put under the executive charge of General Leslie Groves. Oppenheimer was made responsible for the design and construction of the bomb, and appointed scientific director of a vast new laboratory town that was quickly built at Los Alamos, with groups of scientists and engineers working at other centers on different aspects of the whole program.

Teller joined Oppenheimer at Los Alamos, but Szilard remained with Fermi in Chicago to build a reactor to produce the artificial fissile element, plutonium. In 1944, Fermi, who had helped plan the work that had to be carried out at Los Alamos, was appointed director of the independent Argonne Laboratory, an offshoot of the Chicago operation. Oppenheimer had wanted Rabi to be his associate director, but his offer came too late. Rabi was already at work with Lee DuBridge in the central radiation laboratory that had been established at MIT to develop microwave radar. Oppenheimer had to be content with his help as a personal adviser and consultant.

It is clear from the histories and biographies of the period that at the start there were no disagreements about what needed to be done to get the nuclear project moving, and about the direction in which it should be steered. What was known about nuclear physics pointed to two possibilities: either the atom (fission) bomb or the vastly more destructive hydrogen (thermonuclear or fusion) bomb. Both demanded highly sophisticated development work from known principles, but because the hydrogen bomb would need to be triggered by a fission device, it was obvious that the latter, i.e., the simpler atomic bomb, would have to be the first objective of the Los Alamos laboratory, which rapidly grew into an organization in which thousands of scientists, engineers, and technicians were at work. “On an intellectual level,” said Teller of his boss, “in following the thousand things that were going on in the laboratory, understanding them all, influencing them all for the better, he did a truly magnificent job. Furthermore, he knew the character of every one of the people and he used his knowledge to get effects, to influence people.” To Teller, however, the way that Oppenheimer ran the organization was “deeply repulsive.”7

The use of the words “deeply repulsive” after the lapse of seventeen years from the events about which he himself wrote in his autobiography is truly remarkable. Teller, so his biographers write, was a malcontent from the moment he joined the Los Alamos laboratory. He clearly felt that he was not appreciated as much as he should have been. He had taken an active part in the discussions that laid out the program of work of the theoretical division of Los Alamos, and had agreed how the work should be distributed. He took it as an affront that Bethe, not he, was then made head of the division. According to Bethe, he regarded himself as having seniority over everyone else at Los Alamos, including Oppenheimer. He was not prepared to play second or third fiddle. Bethe recognized that if the Los Alamos project was to succeed, it was necessary to adhere to decisions that had been taken about the detailed calculations that had to be made, and that much administrative work would be inevitable. Teller on the other hand thought that Bethe overorganized the work.8 He did not want to become “part of an organization.”9 “Subtle differences,” Teller also says, soon started to develop between him and Oppenheimer,10 despite the fact that he had to recognize that Oppenheimer knew “how to lead powerfully without seeming to do so,”11 and that he had established himself as the intellectual leader, not simply the scientific director, of the whole enterprise.

Teller writes that while he agreed that the fission bomb had to be the first priority, Oppenheimer and Fermi “urged” him as a theoretical physicist to devote his time to the longer-term calculations that would define the specifications of a fusion bomb.

It would seem, however, that he resented the fact that as the work on the fusion bomb progressed he did not see as much of Oppenheimer, and indeed of Bethe, as he felt entitled to do. According to Bethe, in the end Oppenheimer had to set aside an hour a week to talk to Teller—partly, no doubt, to soothe the latter’s amour-propre. One day when Bethe asked Teller to turn his attention to an extremely important problem relating to the implosion system of a warhead, which to Bethe was “the most important task in the theoretical division,” he was met with a refusal.12 Teller told his biographers that he did not recall any such incident, although he did admit that he and Bethe did not work well together. One would judge that Bethe’s memory was the better, since Oppenheimer had to write to General Groves to say that since Teller was unsuited to the responsibility of carrying out the work that he had been assigned under Bethe, he was being replaced by Rudolph Peierls.13

The next significant occasion when Teller found himself at odds with Oppenheimer was shortly before the Alamogordo test shot in 1945, when Szilard asked him to canvass for signer’s among the Los Alamos scientists to a petition that had been drafted by the Chicago reactor team. It called on the United States not to use the atomic bomb without first warning the enemy. Teller has written that he was in favor of signing, but that he was dissuaded by Oppenheimer, who told him that “he thought it improper for a scientist to use his prestige as a platform for political pronouncements,” and that the matter was being considered by the responsible authorities in Washington.14

In the autobiography that he published seventeen years later, Teller wrote that he regretted that he did not circulate Szilard’s petition. With hindsight he argues that the “tragic surprise bombing” of Hiroshima “was not necessary since a nighttime atomic explosion high over Tokyo…would have been just as terrifying as Hiroshima.” In public debate in 1962, the same year that his autobiography was published, Teller declared that he had agreed “completely” with Szilard “that we should not drop the bomb on Hiroshima.” Since Teller was constantly at odds with Oppenheimer, it is difficult to see why he deferred to him on this occasion, and why he did not sign, if not circulate, the Szilard petition—or at least make his views publicly known when they might have carried more conviction.

Teller had rapidly made himself a kind of lone dog so far as the main direction of the Los Alamos work was concerned, and none of the kudos for the successful development of the atom bomb was to fall on his shoulders. Not surprisingly, when the war ended, a large proportion of the Los Alamos team left for home. Teller remained for a few months, hoping that the establishment would now press on not only with the further development of fission bombs, but also with the fusion bomb—“the super.” But Washington had not yet decided what was to be done with Los Alamos. Oppenheimer had handed over the directorship to Norris Bradbury, an amiable and able Navy scientist, and Bradbury had to tell Teller that what he wanted was unrealistic.15 Teller accordingly returned to Chicago, thoroughly dissatisfied with the way things were going, and determined to urge the claims of “the super” from the sidelines.

When in 1949 news of the first Russian test of an atomic bomb was broadcast, Teller picked up the telephone and called Oppenheimer, who by then had been appointed director of the Institute for Advanced Study at Princeton, and also chairman of the General Advisory Committee of the new Atomic Energy Commission. “What do we do now?” he asked. The reply, as he records it, was “keep your shirt on.”16 That was not what Teller wanted to hear. But he got the support of Ernest Lawrence, the distinguished Berkeley physicist, and intensified his lobbying campaign for the super, finding sympathetic listeners both in Senator Brien McMahon, the chairman of the Joint Congressional Committee on Atomic Energy, and in Lewis Strauss, a member of the AEC. The Joint Chiefs of Staff were also in favor of developing the superbomb. Heavy pressure was being brought to bear on President Truman to give the go-ahead.

On the other hand, Oppenheimer’s General Advisory Committee unanimously advised against embarking on the new development. In their view, fission bombs were powerful enough to serve as a deterrent to war. Rabi and Fermi, who were members of the committee, went further, and urged the President to call an international meeting to seek world-wide agreement that no country would ever embark on the development of the super, a weapon that would have unlimited destructive power. Fermi and Rabi wrote of the H-bomb, “It is necessarily an evil thing considered in any light.”17 Truman decided otherwise. In 1951 Los Alamos was given the order to press on vigorously. As members of the General Advisory Committee, Oppenheimer, Rabi, and Fermi were kept in touch with the progress of the work, without being directly involved as Teller was.18

This should have been Teller’s moment. He had returned to Los Alamos for a year’s work a short time before the Russian test. There were no doubts about the principles that underlay the idea of the fusion bomb that he had in mind, but his sums were not coming out as he had hoped, and he was finding it extremely difficult to produce a design that would actually work. He was also at odds with Bradbury and with everyone else who mattered. Apart from Lawrence and two or three others in the upper echelons of the nuclear weapons fraternity, he simply could not get on with his peers. Soon he was campaigning, with the help of Lawrence, for the establishment of a second nuclear weapons laboratory. Friendly competition, he proclaimed, is a good thing. Whether or not that was the real reason why he wanted his own show, the Lawrence Livermore Laboratory, established in 1952, was to give him the chance to surround himself with his own band of young acolytes. He was not offered the directorship of the new establishment but, that apart, he could no longer blame his frustrations on Oppenheimer, Bethe, Bradbury, and all who thought like them.

Two years later, in April of 1954, came the event that not only perpetuated Teller’s isolation from the leaders of American physics, but also led to what he bitterly regarded as his ostracism by the general scientific community.

Oppenheimer had by then ceased to be a member of the General Advisory Committee, but was still one of its consultants. Teller had found a receptive audience in right-wing quarters for his complaints about Oppenheimer. It was the era of McCarthy. The FBI started to dig into Oppenheimer’s prewar left-wing connections. The upshot was that he was publicly arraigned by the Atomic Energy Commission as a security risk. Teller was among those who testified at the inquiry. He also provided written testimony to what would have been a criminal charge against Oppenheimer “for having lied to a security person”19 had it not been for the statute of limitations. More significantly, he had made it plain to an official of the Atomic Energy Commission that he wanted Oppenheimer “unfrocked” whatever the outcome of the hearings, otherwise, so he said, scientists would “lose their enthusiasm for the program.”20 In the hearing he declared that he had “always assumed” that Oppenheimer was loyal to the United States, but added that he would “like to see the vital interests” of the United States “in hands which I understand better and therefore trust more.”21

To those who instinctively believed like Teller that greater destructive power meant greater military strength, anyone who opposed the super had become an enemy of the state, and accordingly was not to be trusted. To them, whatever his wartime and postwar achievements, Oppenheimer was a security risk. In the face of this prejudice, an abundance of distinguished and powerful testimony on Oppenheimer’s behalf was impotent. His public career had been doomed from the moment the AEC decided to bring the action. There had been no need for Teller publicly to add his nail to the coffin. Presumably his antipathy toward and jealousy of Oppenheimer impelled him to seek the twisted satisfaction of putting in a personal appearance.

The battle for the super had gone the way Teller had wanted, and the second weapons laboratory for which he had campaigned so hard was authorized. According to his biographers, Teller intervened in the Oppenheimer case out of a sense of “duty.”22 Yet in the letter about the Chicago petition which he had sent Szilard only a few years earlier, he had said that he had worked on the bomb not from a sense of duty, but because the problems interested him—“a sense of duty could keep me out of such work.” The word “duty” clearly enjoys a variable meaning in Teller’s vocabulary.

Rabi was one of the prominent scientists who testified on Oppenheimer’s behalf. In his view, whatever Oppenheimer’s political contacts during the years that led up to the war, however arrogant he appeared to some, it was nonsense to brand him a security risk. Like most other scientists who had been involved in the Manhattan Project, Rabi was also contemptuous of Teller for the part, direct or indirect, that he had played in precipitating the action. He was later to say that he had never known Teller to “take a position where there was the slightest chance in the interest of peace. I think he is an enemy of humanity.”23

Szilard played no part in the hearings, but judging from the record of his two debates with Teller, he would have had good reason to echo Rabi’s rebuke. Teller opposed Szilard on every major issue that the latter felt was significant if the nuclear arms race was not to end in tragedy.24 Unlike Szilard, he saw the Russians as being ready to embark on nuclear war in their quest for world domination. They cheated and in his view could never be trusted to keep any arms control agreement. The only way to deal with them was to build a more powerful nuclear arsenal than theirs, and to see to America’s defenses. Szilard was wasting his time in trying to promote US-USSR accord. The Teller who debated with his old mentor had become a black-and-white propagandist in favor of the cold war.

The irony is that only a few years before, he had made speeches and written papers in which the views he expressed were close to those that were the basis of Szilard’s crusade for peace. In 1946 he had written that now that nuclear weapons had been invented, nothing but world union could provide a safe environment for the next generation. 25 An atomic war could endanger the survival of mankind, and despite the Russian rejection of the Baruch plan for the control of nuclear power by the United Nations, the US should strive to gain their cooperation. “World Government,” Teller wrote, “is our only hope for survival.” The old Teller might have been Szilard in disguise. It is difficult to avoid the thought that the switch in Teller’s views was a matter of opportunism, an effort to adjust what he was saying to the opinions of his new right-wing patrons.

For the new Teller, champion of the “super,” had become something of a hero among hard-liners, and also the subject of much publicity, even if, reciprocally, a pariah in the social circle he had frequented before—the small and closely knit group of theoretical and experimental nuclear physicists. At first Teller did not try to disguise the extent to which he was bruised by the fact that “more than ninety percent”26 of the scientific community now regarded him as an outcast. He wanted to issue a public statement saying that he had not meant what everybody had understood his words about Oppenheimer to mean. He was dissuaded from doing so by Lewis Strauss, the member of the AEC into whose ears he had long been pouring his complaints about Oppenheimer.

In 1955 in an apparent effort to repair the damage he had done himself, Teller published an article headed “The Work of Many People,” in which, without disavowing his public image as the “father of the hydrogen bomb,” he “lavishly praised” the contributions made by other physicists and mathematicians, including Oppenheimer, Bradbury, and Bethe.27 Among others to whom he also referred admiringly were Stanislaw Ulam and Frederic de Hoffmann. He wrote that in the final days leading up to a decision about the correct configuration of a practicable fusion warhead, an “imaginative suggestion” was made by Ulam, and a “fine calculation” by Hoffmann.

The story then takes a curious turn. In the preceding year (1954) James Shepley and Clay Blair, Jr., two journalists, had published an account of the genesis of the bomb which, in Teller’s words, had praised him to high heaven at the same time as it had denied others some of the credit that should have fallen on their shoulders. According to these two partisans of Teller, Ulam’s “imaginative suggestion” had stimulated an idea in Teller’s mind which he immediately passed on to Hoffmann—described by Blumberg and Owen as Teller’s “right hand man”—who then set about his fine calculations.28

But Teller soon started to deny that he had been stimulated by anything that Ulam had said. “Ulam triggered nothing,” he told his biographers.29 The idea was his own; it occurred to him just before the conversation he had had with Hoffmann soon after he had spoken to Ulam. In his autobiography, published in 1962, Teller had written that Hoffmann not only made “a fine calculation,” but also a “projection” of the idea, so revealing how an effective thermonuclear bomb could be constructed.30 Hoffmann then wrote up the results of his work in an important paper in which he not only omitted any reference to himself, but which he signed with Teller’s name. In 1962, Teller had written that he felt “ashamed” that he had allowed this to be done. In his new book, published last year, he no longer feels “ashamed”—he is only “sorry.”31 But nowhere did he explain how one man could come to sign another’s name to a piece of work that the latter had not written. Stanislaw Ulam was clearly not as docile as Hoffmann. He and Teller have never made up their quarrel about their respective contributions to the critical concept in the ultimate and successful design of the super.

Teller’s efforts to make amends for his behavior toward Oppenheimer and Ulam got him nowhere. Instead of enjoying the recognition that would have been accorded a scientist who had made significant contributions to the body of scientific knowledge, he has had to rest content with the admiration of those who view the future as a continuing conflict between East and West. He was to develop into an ally if not a leader of the “hard right.” Whether he ever appreciated that this would happen is unclear, but it was not long before he was writing and speechifying like the best of them. The higher he flew with the hawks, the more it seemed to compensate for his ostracism by those whom he had regarded as his peers—and the more he was impelled to justify and rationalize his actions.

Thus, where before he had feared that a nuclear war would endanger the survival of the human species, he now declared that through the use of “clean” warheads a nuclear conflict could be confined.32 In 1967 he saw a most menacing and special danger coming from China, which he thought would be able to blackmail the American government by threatening to efface the “12 biggest cities in the US.”33 Nuclear testing had to go on because new warheads had to be designed. A test ban agreement would be a disaster because the Russians could not be trusted. Radioactive fallout was not as dangerous as was usually assumed. The Teller who could with feeling write after Hiroshima of “the picture of fires raging unopposed, wounds remaining unattended, sick men killing themselves with the exertions of helping their fellows”34 would become transformed into a Teller who could reassure Europeans not to worry because even if a nuclear war would wipe out millions, some of them would still survive.35 In one of their public debates, he told a highly skeptical Szilard to accept as a “concrete proposal” the proposition that Atlantic union under one government should be made a political reality.36 Szilard, romantic though he was, emerges from their television exchanges as by far the greater realist. That was more than twenty-five years ago. The Atlantic union under a single government, which Teller saw as a concrete proposal, is as far away today as it was then.

A European comparing Szilard’s book with Teller’s might well say that Teller sacrificed a large part of his intellectual freedom when he tied himself to the instinctive anti-Communist right. The more his new book is studied, the more pointless it becomes to try to separate fact from exaggeration, and fancy from reality. One has either to swallow Teller whole or to be on the constant lookout for statements which—to use a phrase recently made memorable in the UK—“are economical with the truth.” One runs into difficulties straightaway. The sword in his title is the nuclear bomb; the shield, President Reagan’s mythical SDI space defense system. The opening chapter is entitled “Three Controversies.” The first is supposed to have related to the making of the atomic bomb; the second to the technical and strategic arguments that surrounded the development of the hydrogen bomb; and the third to the current debate about the wisdom of embarking on the development and presumably eventual deployment of SDI—should it ever materialize.

It is farcical to equate these three issues. No controversy surrounded the making of the atom bomb; first, because it was developed in such dense secrecy that not even the vice-president of the US or the senior members of Winston Churchill’s cabinet were aware that there was anything to argue about and, second, because it was developed at a time when the general view was that greater destructive power necessarily implies greater military strength. The controversy provoked by the hydrogen bomb was of a totally different kind from that which surrounds SDI. The central questions about the fusion bomb were: first, whether there was a need for more destructive power than could be encompassed by atom bombs; and second, whether before proceeding, the US should try to negotiate with the Russians a ban on its development. There was also a technical controversy that involved a relatively small number of people—how to design the bomb.

The debate about SDI is altogether different. Quite apart from the momentous technical and financial issues that have never been resolved, there has to be a sensible answer to the question whether SDI is advantageous or disadvantageous from the strategic point of view. This consideration Teller does not discuss, or at best treats as a play of words. He implies that shields and swords are mutually exclusive items of armament. This they certainly are not. They are always complementary. The question is not merely whether a shield is adequate to counter an opposing sword, for example whether in the design of a tank the armor plate can withstand antitank fire, but whether—to pursue that illustration—the penalty that the weight of the armor imposes is acceptable from the point of view of the speed and firepower of the tank.

In 1962 Teller seems to have recognized this fact—an ABM defense, he wrote, is “really good only if it is not much more expensive than the offense,” and provided that it cannot be outwitted with ease37—in Paul Nitze’s updated terminology, only if it is “cost—effective at the margin.” In his new book, Teller conveys the impression that while he agrees that it is unlikely that a perfect ABM defensive system could be devised, the effort to build one would be worthwhile because “the existence of defense at least will save lives.”38 Yet at the same time he admits that more defense will stimulate “additional aggressive deployment,” so that “work on armaments would proceed indefinitely.” But having stated the obvious, Teller then brushes it aside with the unsubstantiated non sequitur that “defense has the advantage that it can provide the basis for a decrease in tensions.”

To write this in the face of the Russian reaction to SDI is truly extraordinary.39 President Reagan told the world that the aim of his Strategic Defense Initiative is to make nuclear weapons “impotent and obsolete.” There are many good reasons to doubt that he has American weapons in mind. These weapons could be made impotent far more cheaply than by the SDI program. Teller, who has not disavowed the widespread belief that he provided the President with the material for his SDI dream, wrote in 1962 that “it would be wonderful if we could shoot down approaching missiles before they could destroy a target in the United States.”40 Surely it was the same missiles to which he was referring in 1982 or 1983, or whenever it was that he bewitched the President with talk of a space-based defense system. Whatever else, and whether or not SDI spurs the Russians to devise countermeasures, it will certainly not encourage them to make “deep cuts” in their ICBM arsenal. One side is hardly likely to help the other in a process by which it might be disarmed. Teller simply deludes himself—not the Russians, or those Europeans who are looking on from the sidelines—when he says that SDI is a basis for reducing tensions.

I keep asking myself whether he really believes what he writes about SDI. He says that he wants to discover whether a space-based ABM system is feasible. “My argument,” he writes, “is for knowledge and against ignorance…. Scientists have the responsibility to make knowledge available and to explain its possible applications”—not to say how the knowledge, once acquired, should be used.41 It is for the people to decide. This message, with the word “duty” sometimes replacing “responsibility,” echoes in Teller’s writings. The proposition seems so disingenuous that one wonders whether it represents more than post-hoc rationalization and self-justification. Obviously the dictum is true for men and women with science degrees who take on jobs with contractual responsibilities, which they then have a “duty” to try to discharge. But scientists have no obligation in general to “make knowledge available.” The few who are competent to add to the store of basic knowledge are not discharging any “duty.” Like composers or painters, they are creative people who are doing what they want to do, doing what they believe they can do, and what they believe is worth doing.

The men who formulated the design specifications of nuclear warheads did not invent the “principles” of nuclear physics. It was the other way around. The specifications were derived from basic principles about the atom that had already been established. If one accepted the Teller dictum, it could be argued that competent chemists who know how to do the job have the “responsibility,” the “duty,” to invent, say, a noxious gas so many thousand times more toxic than any yet known that given the fracture of a single cylinder, the gas that escaped could kill, for example, the population of San Francisco. The gas having been invented, Teller would have it that it would be up to “the entire community” to decide whether it should be used. To Teller, that is “the main principle on which a democratic society rests.” As one reads what he writes, one may well ask whether he appreciates what the words “duty” and “principle” conventionally imply.

In the final exchange of his 1962 debates with Teller, the transcripts of which are reprinted in Toward a Livable World, Szilard made the point that “following the line of least resistance” is a characteristic of the American political system, and that in talks that he had given at ten universities and colleges, he had asked the students whether instead of this deep-seated weakness of the system, a minority could be found that would formulate “a set of political objectives on which reasonable people can agree.” In his reply, Teller said that “reasonable people are people all across the country. No minority should agree, all of us must agree, and that is the way how a democracy has to work.”

Szilard: “All of us cannot agree.”

Teller: “I think all of us will.”42

Clearly in Teller’s mind no minority should disagree. This is a strange statement since Teller, almost always in a minority, never agreed. That conception of democracy, we should note, is offered by a man who fled from Hitler’s Europe and who tells us that he had no interest in politics before, as it was to turn out, he started associating with people who instinctively encouraged his obsession to develop a “super.” What Teller was in effect saying was that a minority did not have the right to disagree—a precept that clearly had not governed his relations with Oppenheimer and Bethe and the vast majority of his other colleagues at Los Alamos. Small wonder that in a statement about Teller that Szilard drafted in 1963, a year before he died,43 he observed that he found himself disagreeing more with Teller’s attitude to the USSR than he did with that of the US administration.

In his campaign for peace, Szilard—to whom, strange to say, Teller dedicated his book—never compromised the mission he had set himself, despite the fact that, as Teller saw it, this made him unpopular “among those who had power or influence”44—presumably in political or military circles. That was not a risk that Teller ran. If in pursuing his path to the super he had to sacrifice the esteem of his peers, he could at least then enjoy that of acolytes he gathered around him, and of others who had a vested interest in the outcome of his and their labors.

Rabi, who never stopped wondering how his fellow citizens could be so misguided as to allow Teller doctrines to flourish, had always known that the destruction of war could never solve the political problems which beset a world that has been transformed by the technological developments of the postwar years. The pursuit of science was to Rabi a unifying force in the peaceful development of society. The Brookhaven Nuclear Laboratory in Long Island, where universities cooperate in research, and the European Centre for Nuclear Research (CERN) near Geneva, are monuments to his success in making a reality of his beliefs. His mind was always wide open to new and challenging ideas, whereas to the public that of Teller appears to have become more and more closed, with his attitudes becoming as rigid as the anti-Semitic prejudices that enclosed him in a gilded ghetto in his youth, and as enduring as those that characterized the ghetto from which Rabi’s family fled. In escaping from the ghettos that they knew, Rabi and Szilard discovered how to enjoy true individual freedom. Teller, who seems to have created a new ghetto for himself, has still to show that he understands what it is that his erstwhile friends discovered.

This Issue

March 31, 1988