“Blackett as Scientific Leader: Physics, War and Politics in the Twentieth Century” full text speech: Mary Jo Nye, Oregon State University, at Imperial College, 26 January 2005

 

lecture poster

Ladies and gentlemen, friends and colleagues, it is a privilege and a pleasure to speak before you today about one of the distinguished scientists of this institution, Patrick Maynard Stuart Blackett, for whom the Blackett Laboratory is named. I am grateful to the Department of Physics, especially Professor Sir Peter Knight and Dr. Martin McCall, for hosting my visit along with the Harvard University Press, especially Ann Sexsmith and Fiona Wyatt. I also would like to say how pleased I am to thank publicly Patrick Blackett's daughter Giovanna Bloor for her generous help and friendship as I have been writing my book on Blackett's scientific life and work. One of the great surprises when I embarked upon this book was that Blackett had not been the subject of a recent biography. The reason for this lapse is most likely the commanding presence of the excellent book-length biographical obituary written by Sir Bernard Lovell in 1975 for the Royal Society. My own biography is different, but certainly not better. We have available now, too, a centenary volume of essays on Blackett edited by Captain Peter Hore of the Royal Navy, following symposia marking Blackett's birth in 1897.

 

Some time ago Giovanna Bloor asked me what drew my interest to her father. The simplest answer to this question is that my work as a historian has concentrated on the history of the modern physical sciences, and Blackett is one of the most important experimental physicists of the twentieth century. However, that simple answer still begs the question.

Two aspects of Blackett's career captured my attention originally. One is the elegance of the narrative history of Blackett's bold hypothesis in 1947 that the Earth's magnetic field can be explained in a simple fundamental equation, having to do with the Earth's rotating mass, and Blackett's own subsequent disproof of his hypothesis using an improved magnetometer of his own design. This scientific case-history has the ethical character of one of Aesop's fables and the philosophical virtue of supporting Karl Popper's demarcation of science from non-science on the basis of falsifiable hypotheses. What appeared to be Blackett's failure in proving his hypothesis alsoo turned out to be the pathway to Blackett's and other scientists' revolutionary studies of rock magnetism and evidence for the old hypothesis of continental drift. I talked about this aspect of Blackett's work at the Blackett symposium here at Imperial College in April 1998.

The reason for my writing a full-scale biographical study of Blackett is broader in scope. I was struck by the controversial life that Blackett, the physicist, chose to lead in war and politics. This is the Blackett who was said by The Times at his death in 1974 to have been a “Radical Nobel-Prize Winning Physicist” who had been “committed too far to the left for [even] a Labour Government to employ with ease.”  While many regarded Blackett as a hero for his achievements as a British physicist and his wartime role in operational research, others vilified him for his postwar criticism of British wartime and Cold War military strategy.

In this criticism Blackett publicly questioned the Allied wartime bombing of civilian urban centers, and he denounced postwar discussions of the tactics of mass destruction as a normal operation of war. Blackett became the first person to openly argue that the United States had used the atomic bomb in Japan "not so much as the last military act of the Second World War, as the first act of the cold diplomatic war with Russia."  Outraged Americans characterized Blackett's statements and his opposition to their development of atomic weapons as a Stalinist apology full of political prejudices. George Orwell in 1949 included Blackett on a blacklist of thirty-eight crypto-communists or fellow-travellers that Orwell drew up for the British Foreign Office.

Blackett's politics were leftwing, rooted in Fabian socialism. He was part of the Soho dining group in the 1930 called Tots and Quots, which included the scientists Solly Zuckerman, Desmond Bernal, Julian Huxley, Joseph Needham, C. H. Waddington, C. D. Darlington and Lancelot Hogben, and the science journalists J. G. Crowther and Peter Ritchie Calder.  By the late 1950s some of this group were meeting at the Reform Club or Brown's Hotel with Hugh Gaitskell, Harold Wilson, Dick Crossman, and other members of the Labour opposition. In his politics, as in his studies of physics or of the operations of war, Blackett demonstrated absolute confidence in the power of rational scientific thinking to solve problems. He made misjudgements and mistakes, of course, and he was too sanguine, or ill-informed, about the bloody politics internal to the Soviet Union, but he displayed stubborn courage in choosing to expose himself to public debate while still pursuing an active scientific life of research, teaching, and administration.

What, we might ask, were some of the historical circumstances that drove Patrick Blackett into the scientific and political life that he chose? What were the personal characteristics that enabled him to achieve a status that his friend, the geophysicist, Teddy Bullard described as that of “the most versatile physicist of his generation”? Finally, what were Blackett's own notions of leadership, and how did he become one of the most powerful scientific figures of his generation? In what follows, I will briefly review aspects of Blackett's education and career, and then comment on ways in which he was perceived by his contemporaries. I finally will turn to the theme of scientific leadership, with remarks about some other scientists whom Blackett admired, concluding with observations on the character and significance of Blackett's role as a scientific leader in the twentieth century.

Building Character, Learning the Trade

 

Patrick Blackett arrived at Cambridge and the Cavendish Laboratory in 1919, and he arrived in uniform. Born in London, Blackett was a middle child with two accomplished sisters. He remembered his childhood as one “brought up in the kindly security of an Edwardian middle-class home.”  It was a home in which overt affection was not shown, and children were not praised, “lest they became conceited.”  As a boy, Blackett spent a good deal of time constructing wireless sets and model aeroplanes. Just before his thirteenth birthday, in September 1910, he entered Osborne Naval College, joining an elite group that included the future King George VI.

 

At that time, in 1910, admission to Osborne was ranked at the prestige of a Winchester scholarship, and the naval schools probably provided the best science and engineering education available in any secondary school in Britain.  In addition to exercises in laboratories, all cadets were expected to learn the rudiments of using tools in pattern-making, fitting, and turning and forging. They operated lath es and engaged in metal filing, as well as in carpentry. Mathematics and modern languages were well taught, as was naval history.

Leadership was regarded as a natural attribute of character. Admission boards were explicit about the qualities they sought:

What is the right sort of boy? . . . that boy has the best chance [who] is resourceful, resolute, quick to decide, and ready to act on his decision. He must be no slacker, but keen to work and play. He should be sound alike in wind and limb, and in the big and little principles of conduct . . . . He should give promise of being responsive and observant, closely in touch with his surroundings, but master of himself. The boy of sensitive, poetic spirit, the ruminating young philosopher, the scholar whose whole heart [is] in his books are types that have a real use in the world, but their proper place is not the Navy.

In some sporting activities, Blackett was not as successful as other boys. Captain Lord Alastair Graham later read to Blackett from notes that he had made on Blackett as a cadet: “Games: does not shine.” Still, part of Navy lore is that Blackett was kicking off at a students' football match when he was informed of winning the Nobel Prize.  As a naval cadet at Dartmouth, Blackett enjoyed sailing in the Estuary. He was an avid birdwatcher, and he took photographs with a camera that he built himself, a prelude to cloud-chamber photographs that would later make him world famous as a physicist.

On August 1 st , 1914, when war broke out, 400 cadets at Dartmouth were told to pack their chests. Blackett saw action in the Falklands in 1914 and at Jutland in 1916. He witnessed the battle cruiser HMS QUEEN MARY sink, and thousands drown. His own battleship HMS BARHAM took fire in the battle, with twenty-four deaths on board. As First Lieutenant controlling gunnery fire on HMS Sturgeon , Blackett's last action was a destroyer battle off Terschelling (Netherlands) in April 1918. He was in charge of controlling shell damages to the ship and getting it back to Harwich. He managed to get off some 50 rounds of gunnery shots as well.

When the war was over in November 1918, the Admiralty decided to send some 400 junior officers to a six-month course of general studies at Cambridge University. Still in uniform, Blackett dined for the first time at Magdalene College in late January 1919. A few days later, he wandered into the Cavendish Laboratory. Three weeks later he resigned from the Navy. By May 1921 he received his degree, with a Second Class in the Mathematics Tripos and a First in Part II of the Physics Tripos.  He was one of the few physicists of his generation to have served and survived combat in the war before completing his university studies. He was unusual among fellow university science graduates in his self-discipline, self-reliance, and experience of leadership.

Ivor Richards, a young philosophy don at Magdalene College, and later professor of literature at Harvard, recalled first meeting Blackett in 1920.

I was living . . . in a garret in Free School Lane a door or two from Cavendish Laboratory. 'Came a quick step on its ‘rotten-runged, rat-riddled stair,' a tap on the door, and there entered a young Oedipus. Tall, slim, beautifully balanced and looking always better dressed than anyone. People used to ask him the name of his tailor . . . . [Above] was that mysterious intense and haunted visage, which later made [Jacob] Epstein count this Nobel Prize winner's bust among his greatest . . . .

Many descriptions of Blackett as physicist remark on his appearance, including his height at six feet two and one-half inches. Some commentators remark that his looks were somewhat misleading. “Tall and strikingly handsome . . . he might have been predominantly formidable. But he was not. Besides being humane, he was witty, amusing and very good company.”  Of Blackett in his 40s, a former student recalled “He was tall and strikingly handsome.”  Of Blackett in his 50s, the French newspaper France Soir suggested he could have been a cinema star. Of Blackett in his 60s, at Imperial College in London, a friend wrote: “Even more handsome as an aging man than as a young one, he was as much a figure in the King's Road, as he had once been in the King's Parade.”

In March 1924 Blackett married Costanza Bayon, a student of modern languages at Newnham College. She had an Italian father, but had been raised by an English couple in Rome who nicknamed her “Pat.”  Thus, to intimate friends, they were the two Pats. During the 1924-1925 academic year, they lived in Göttingen, where Blackett worked with James Franck in electron physics.  Returning to Cambridge, the couple settled at 59 Bateman Street, where they kept open house at least once a week. Reports were that their guests tended to be semi-bohemian and left-wing. Ivor Richards called them the “handsomest, gayest, happiest pair in Cambridge.”

From 1921 to 1933, Rutherford was Blackett's first and only research director. Assigned by Rutherford in 1921 to modify an automatic cloud chamber for the study of alpha particles bombarding targets, Blackett worked for several years to perfect the instrument in the face of Rutherford's impatience for results.  Rutherford had assigned Blackett the problem of confirming Rutherford's hypothesis that alpha particles incident upon nitrogen gas induce a transformation in the nitrogen nucleus, with expulsion of a hydrogen particle (proton) from the nucleus. In the summer of 1924, Blackett obtained eights tracks confirming rearrangement in a nucleus. These photographs have been widely reprinted ever since.  The work made Blackett's reputation at the age of twenty-seven.

Blackett's second set of really famous experi ments came in the early 1930s when he collaborated with Giuseppe Occhialini on devising a cloud chamber in which expansion of the cloud chamber is triggered by pas sage of charged particles in cosmic radiation. While the two were accumulating data and discussing its theoretical implications with Paul Dirac in fall 1932, C arl Anderson at Caltech announced his discovery of a positively charged electron in the cosmic radiation. Anderson initially characterized the particle's production as a rare event. In contrast, Blackett and Occhialini used their data to explicitly link the anti-electron to Dirac's relativistic electrodynamics, a theoretical insight that had not occurred to Anderson.

Although Blackett and Occhialini were nominated immediately for a Nobel Prize for their work, it was Anderson who received a share of the Nobel Prize in Physics in 1936 for his “discovery of the positron,” along with Viktor Hess who had established the existence of cosmic radiation. No doubt, Blackett and Occhialin i were disappointed in this outcome. Blackett also was chagrined by Rutherford's cool response to the positron work when Rutherford told Blackett that he would prefer to see the positron particle produced in radioactivity transformations rather than in cosmic radiations.

1933 turned out to be a pivotal year. When Rutherford told Blackett that he thought Blackett should spend more time teaching and that he, Rutherford, had no intention to expand facilities for nuclear physics at the Cavendish, Blackett left Rutherford's office white-faced with rage. Outside the door, Blackett told his postgraduate research student Frank Champollion that “If physics laboratories have to be run dictatorially . . . I would rather be my own dictator.”

Blackett often was asked about the Cavendish years and about the legendary Rutherford. Blackett praised Rutherford's power of concentration and pictorial imagination, his eye for the unexpected, and his boundless enthusiasm, saying that he, Blackett, had “learnt early [from Rutherford] the vital importance of the role of the director of research in selecting promising problems for his research students.”  This would be Blackett's method, too, for directing a research school.

A Lab of His Own

In 1933, Blackett left the Cavendish Laboratory for Birkbeck College in London, where he took charge of the physics department and laboratory. In 1937 he succeeded Lawrence Bragg in the physics chair at Manchester, a position which previously had been held by Rutherford. And, of course, Blackett would return to London in 1954, to Imperial College, taking many of his Manchester laboratory colleagues with him. Shortly after he arrived in London, in 1933, Blackett was recruited by Henry Tizard, in early 1935, to join an Air Ministry committee charged with investigating the use of radio waves in air defense. Although many members of the British Left were taking strongly pacifist positions in the mid-1930s, Blackett parted company with them in this matter. He was no pacifist. By 1940 Blackett became scientific adviser to the Army's anti-aircraft command, organizing a group of scientists to study the operational use of radar sets, guns, and mechanical calculators for anti-aircraft fire. [Edward Bruce] Hamley's classic nineteenth-century textbook The Operations of War (1867) would be updated now by methods of scientific analysis.

In the Royal Air Force's Coastal Command, Blackett headed a group that recalculated depth settings for anti-submarine explosives. Moving to the Admiralty in 1942, Blackett directed an operational research group that brought about significant improvement in the use of airborne radar for finding German submarines which were sinking merchant ships in the Atlantic. This work often is credited as a turning point in the war, by summer of 1943, so that American supplies and troops could reach England for the invasion of Europe.  It was this work, too, that brought Blackett into what later became a well-known confrontation with Prime Minister Churchill's scientific advisor Frederick Lindemann and that laid the basis for Blackett's public critique in 1948 of the wholesale bombings of German cities as both inefficacious and immoral.

Taking on increased university and governmental duties after the war, Blackett remained active both in personal research and in directing the physics department at Manchester. He continued to be “hands-on,” showing young Clifford Butler, for example, how to set up the cloud-chamber control mechanism for cosmic-ray experiments at Manchester. He closely followed Butler and George Rochester's experiments that photographed tracks of a new “strange” particle identified by its signature V-track, and Blackett worked with them in writing up the results on what later would be called kaons, although his name did not appear on the published paper.

The military experiences of his youth and his maturity deeply informed Blackett's style of leadership in his laboratories. Images of military bearing and conduct frequently have been used to characterize Blackett: his very presence was said to carry authority. At Manchester, a former student reports that he “was awestruck by [Blackett's] stately procession down the main stairs for lunch. He always walked in the dead centre of the staircase, disdaining the banisters. He held his hands, naval fashion, in his jacket pockets, with thumbs protruding. He had no nickname: he was Professor Blackett.” In an interview with Brian Connell of Anglia Television, Blackett admitted that his laboratory staffs said that he ran a department like a Captain runs a ship: “ There is something I think in the tradition of delegating authority completely to young, junior people and then if things go wrong, taking the blame yourself.” That this style of leadership had a practical effect in research, as well as in management, is noted by Francis Everitt, one of Blackett's postgraduate students at Imperial in the late 1950s. Everitt did his thesis work on rock magnetism and reversals of magnetic polarity in sedimentary and igneous rocks. Two remarks encapsulate Blackett's outlook, says Everitt: “Make sure you gather plenty of data,” and “You should treat your research like a military campaign.”

Everitt himself came to Imperial College as an undergraduate and stayed on for his postgraduate degree. It was precisely because of Blackett's failed experiments on the Earth's magnetism that Everitt wanted to study with Blackett, by his own account. As I mentioned earlier, Blackett's most spectacular failure was his disproof, using his own magnetometer, of his well-publicized hypothesis that the magnetic fields of the sun, stars, and earth are a fundamental property of their rotating mass. A beautifully simple law had related magnetic and angular momentum to the gravitational constant and the speed of light. But the law was wrong, as proven by Blackett and his assistants from 1949 to 1952. Yet Blackett's announcement that experiments had proved his rotating body hypothesis wrong only increased his prestige. Everitt recalls:

By the time I was seventeen, I had heard of Blackett, initially because of the famous cloud chamber photographs on nuclear disintegrations which were reproduced in dozens of books. My physics master in high school . . ., when he was teaching us about magnetism, . . . described Blackett's hypothesis and then explained how, when the evidence told against it, Blackett had demonstrated his scientific integrity by immediately acknowledging that it was wrong. Thus Blackett was held up as a high role model of the physicist by a man who did not in the least share Blackett's political views.

Blackett's personal research at Imperial College had turned to rock magnetism and the use of paleomagnetic data to establish latitude effects for continental drift.

His rock magnetism group included John Clegg from Manchester; and Blackett also brought with him to Imperial other Manchester colleagues and technicians, including Harry Elliot, with a cosmic-ray group, and Clifford Butler, who headed the high-energy nuclear physics group.

When Blackett arrived at Imperial, there were three chairs in the department: Blackett's; David Wright's chair in Technical Optics, and Samuel Devons's Chair in Low-Energy Nuclear Physics. Devons went on to succeed Blackett at Manchester in 1955. By 1960 Imperial had a new Physics building, seven professors, 27 lecturers, and 32 research assistants in the department, divided into ten independent research groups. Students numbered some 300 undergraduates and 100 postgraduates.

As Bernard Lovell has noted, Blackett was one of the first department heads in Great Britain to implement the strategy of multi-professorial departments, aiming to create an urban scientific research and educational institution that equalled the old universities of Cambridge and Oxford, while opening up opportunities for new disciplinary fields and for college students unlikely to enter the older elite establishments. This was a practical application of Blackett's socialist and scientific politics. The commitment led him in the early 1950s to decline to be a candidate for the Cavendish Laboratory directorship, succeeding Lawrence Bragg, or for the Provostship at King's College in Cambridge. The Imperial Physics Department initially had one Nobel laureate in Bl ackett himself. Dennis Gabor, who became Professor of Applied Electron Physics in 1958, received the Nobel Prize in 1971; and Abdus Salam, who became Professor of Theoretical Physics in 1959, after transferring from the Mathematics Department, shared the 1979 Nobel Prize with Sheldon Glasgow and Steven Weinberg.

It is not surprising that Everitt found Blackett to be one of the busiest men he had ever seen in the late 1950s: chairing Royal Society, university, and government committees; a member of Harold Wilson's inner circle of advisers on science and technology policy; a consultant to the Indian government on science and military matters. As a close friend of the Cambridge-educated Indian physicist Homi Bhabha, Blackett also began spending considerable time in India after 1947, where the geophysics group from Imperial set up a paleomagnetic survey of the Indian subcontinent with Indian scientists in the Tata Institute in Bombay. Blackett also became a military and scientific advisor to the Indian.