bryan randell coles. 9 june 1926 −− 24 february 1997...

17
Elected F.R.S. 1991 24 February 1997: -- Bryan Randell Coles. 9 June 1926 David Caplin , 51-66, published 1 November 1999 45 1999 Biogr. Mems Fell. R. Soc. Supplementary data http://rsbm.royalsocietypublishing.org/content/suppl/2009/04/22/45.0.51.DC1 "Data Supplement" Email alerting service here top right-hand corner of the article or click Receive free email alerts when new articles cite this article - sign up in the box at the http://rsbm.royalsocietypublishing.org/subscriptions , go to: Biogr. Mems Fell. R. Soc. To subscribe to on May 20, 2018 http://rsbm.royalsocietypublishing.org/ Downloaded from on May 20, 2018 http://rsbm.royalsocietypublishing.org/ Downloaded from

Upload: haliem

Post on 20-Mar-2018

219 views

Category:

Documents


4 download

TRANSCRIPT

Page 1: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

Elected F.R.S. 1991 24 February 1997:−−Bryan Randell Coles. 9 June 1926

David Caplin

, 51-66, published 1 November 1999451999 Biogr. Mems Fell. R. Soc. 

Supplementary datahttp://rsbm.royalsocietypublishing.org/content/suppl/2009/04/22/45.0.51.DC1"Data Supplement"

Email alerting serviceheretop right-hand corner of the article or click

Receive free email alerts when new articles cite this article - sign up in the box at the

http://rsbm.royalsocietypublishing.org/subscriptions, go to: Biogr. Mems Fell. R. Soc.To subscribe to

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 2: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

BRYAN RANDELL COLES9 June 1926 — 24 February 1997

Biog. Mems Fell. R. Soc. Lond. 45, 51–66 (1999)

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 3: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 4: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

BRYAN RANDELL COLES

9 June 1926 — 24 February 1997

Elected F.R.S. 1991

B D C

Blackett Laboratory, Imperial College of Science, Technology and Medicine,

London SW7 2BZ, UK

E

Bryan Coles’s background and career reflect in microcosm the enormous changes that

occurred in the British educational system in the middle decades of the twentieth century. His

parents, Charles and Olive (née Randell) Coles had both come from families that migrated

from Somerset to Cardiff, attracted by the expansion of the docks in the 1880s, one

grandfather having been a plumber and the other a carpenter. Charles received no formal

education beyond elementary school, despite doing exceedingly well in the scholarship

examination, because his parents were unable to afford his attendance at high school; he

started work as a book-keeper. Olive’s formal education also had been curtailed; she went to

work as a milliner until their marriage in 1923.

Soon after Bryan and his twin sister Isobel were born in 1926, the family was badly

affected by the Depression, and for three years Charles had no full-time employment; he

cycled around Cardiff picking up whatever odd job he could find. Despite these problems,

education remained a high priority for Olive and Charles, and they ensured that all four of

their children received secondary schooling. Bryan’s elder brother Gerald, his younger sister

Brenda and Bryan himself all went on to University College, Cardiff, and Isobel trained as a

nurse at Guy’s Hospital, London.

Olive read poetry to her children, as her father had for her, which must have helped Bryan

to develop his extraordinarily retentive memory, fed later by a voracious appetite for reading.

He learnt to play the flute (purchased with a small legacy), played in the school orchestra, and

conducted the school choir; in adult life this interest was to transform into an enthusiasm for

53 © 1999 The Royal Society

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 5: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

54 Biographical Memoirs

opera. Outside school, he was a keen cyclist, exploring the Glamorgan countryside at

weekends. His first bird-watching notebooks date from the age of ten, the beginning of his

love of countryside, which was later shaped by his discovery of the writings of Gilbert White

and Richard Jefferies.

Bryan attended the local elementary school in Landsdowne Road, close to the family

home in Cardiff, and at age eleven he was awarded a Craddock Wells Scholarship to Canton

High School, the local grammar school. His inclinations were initially much toward the arts

and humanities, influenced no doubt by his parents, but he showed also considerable

mathematical strengths. After sitting his School Certificate, he decided to continue with

mathematics and to take up chemistry and physics; it seems that memories of family

economic hardship were important, and he was unsure how his interests in literature could be

combined with a career. The excellent results of his Higher School Certificate (he achieved the

best mathematics marks in Wales) in 1944 led to an award of a State Scholarship to study

metallurgy at University College, Cardiff: the booming South Wales steel industry seemed to

offer a secure future.

His undergraduate years in bomb-damaged Cardiff came during the austere end-of-war

period, and he took life very seriously. He was somewhat dissatisfied with his metallurgy

courses, as they seemed to shy away from the fundamental understanding that he sought; he

complained later that they had been unable to tell him why copper is red, and so different in

colour from other metals. To satisfy his curiosity, he took himself off to subsidiary physics

and chemistry lectures, and then delved into W. Hume-Rothery’s The structure of metals and

alloys (1936); he had been warned that he would find Mott & Jones’s The theory of the

properties of metals and alloys (1936) too difficult. Bryan graduated in 1947 at the top of his

class with first class honours; it was totally in keeping with his character that, with part of his

graduation prize, he purchased the Mott & Jones text and met the challenge.

O, L USA

Coles wrote to Hume-Rothery, who was then Warren Research Fellow at the Inorganic

Chemistry Laboratory in Oxford, asking if he could start research with him. There could have

been no more propitious choice, for Hume-Rothery had a burning faith in the search for

fundamental understanding. He had absorbed the ideas of Sommerfeld and Bloch, and

promulgated them in his path-breaking text (Hume-Rothery 1931). The Preface set out his

credo:

… the ‘Practical Metallurgist’ has amassed a considerable amount of information in connexion with

the metals of commercial importance, but has made little or no attempt to investigate the properties of

metals and alloys systematically from the point of view of the Periodic Table, in order to discover the

general principles involved. On the other hand, the Physicist in his investigation of the electrical and

other physical properties of metals has not merely often failed to examine these systematically from

the point of view of the Periodic Table, but … much experimental work has been carried out on

specimens which any scientific metallurgist would have condemned as unsuitable.

… the Metallurgist has, in the past, been profoundly suspicious of the value of theoretical work,

…and this very greatly hinders the development of a connected science of Physical Metallurgy.

Hume-Rothery was therefore delighted to find a student with not only a metallurgical

background (there was then no metallurgy department in Oxford) but one whose attitude and

view of the subject lay so close to his own.

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 6: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

Bryan Randell Coles 55

For his DPhil, Bryan investigated the Ni–Mn phase diagram. In classic Hume-Rothery

style, he used the crystal structures, phase compositions and phase boundaries to infer

electronic structures, and particularly the d-shell configurations (1)*. Electrical and magnetic

measurements on these alloys would have been rewarding, and the facilities were available in

the Clarendon, but the interdepartmental barriers were too high to be overcome.

After submitting his thesis in 1950, Bryan was approached by the head of the Physics

Department at Imperial College, Sir George Thomson, F.R.S., and appointed to a lectureship

in metal physics. The appointment was actually a joint one, in both the Physics Department

and the Metallurgy Department, but his teaching was confined to the latter, suggesting that

Thomson deemed the subject not totally appropriate for his own department. Only after

Thomson’s departure in 1952 did a mainstream course on solid-state physics enter the Physics

Department’s syllabus.

On the research side, Bryan began to build up equipment, primarily for the electrical

transport studies (down to liquid nitrogen temperatures only at that time) that he had lacked

in Oxford. Crucially, though, he had access to H. Jones’s (H. Jones, F.R.S. 1952) group in the

Mathematics Department, who were at the cutting edge of theory of the metallic state: E.H.

Sondheimer (transport), S. Raimes (band structure) and E.P. Wohlfarth (magnetism).

For the first two or three years at Imperial College, Bryan had been a one-man band, so

when in 1954 he had the opportunity of an International Co-operation Administration

Fellowship, he leapt at the chance of visiting an energetic and like-minded group working on

magnetism in the transition metals and their alloys at the Carnegie Institute of Technology in

Pittsburgh, Pennsylvania. During his two years there he broadened his experimental

repertoire, acquiring expertise in magnetism studies from A. Arrott and J.E. Goldman, and in

heat capacity measurements from S.A. Friedberg.

Another motive behind Bryan’s visit to the USA was Merivan Robinson, whom he had

met in the summer of 1953, when she was a Fulbright Scholar in London. She returned to a

Fellowship in English at Bryn Mawr College, also in Pennsylvania, but admittedly at the other

side of a rather large State. After a year of commuting weekends, they were married in 1955 in

Merivan’s home city of St Paul, Minnesota.

On their return to London in 1956, they settled into a small flat in South Kensington,

within easy walking distance of the college; Bryan would often return to the department in

the evenings and at weekends. A little later, when property prices in the area rose

astronomically, this proximity was a luxury greatly envied by other Imperial College staff.

After the birth of their two sons, Matthew (in 1965) and Jonathan (in 1967), Bryan and

Merivan moved to an adjacent flat, and after much knocking down of walls obtained a

spacious apartment; here they frequently hosted their wide circle of friends and colleagues

with excellent food, wine and conversation. A few years earlier, they had found a Sussex

cottage where at weekends they could escape the noise of London, and which Bryan could use

to indulge his countryside passions of bird-watching and medieval churches.

The two years in Pittsburgh had seen the formation of a wide range of professional

collaborations, many of which became enduring friendships, so for both personal and

scientific reasons Bryan became an early transatlantic ‘frequent flyer’. He visited the

University of California, San Diego, where he fell under the spell of Berndt Matthias, for six

months in 1962 and again in 1969, and the University of Minnesota for three months in 1983.

* Numbers in this form refer to the bibliography at the end of the text.

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 7: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

56 Biographical Memoirs

Later on, he was a frequent summer visitor to the Los Alamos National Laboratory, attracted

not just by the excellence of their science and the facilities to handle actinides and beryllium,

but also by the opera in Santa Fe and walking in the Bandalier National Monument.

The vitality of American science, and its informality, attracted him. Other Europeans

might find that mode of discussion and scientific argument, sometimes verging on the

aggressive, difficult to deal with, but Bryan could hold his own in robust debate, and deflect

frontal verbal attack with a smile and a disarming response.

In the major redevelopment of Imperial College’s South Kensington site, Physics was the

first department to move into a new building (in whose design Bryan was one of the academic

liaison staff), and so in 1960 more space became available for the Metal Physics Group, and a

new lecturer, J.G. Park, was appointed. Park’s interests in superconductivity added another

dimension to Bryan’s alloy work, and from that time on the two strands of magnetism and

superconductivity were interwoven in the latter’s research. (J. Gavin Park died in mid-career in

1983. His near namesake J.-G. (Je-Geun) Park was Bryan’s last research student (1990–93).)

The post-Robbins era of the 1960s saw a period of rapid expansion of Imperial College in

general, and of Bryan’s group in particular. He was promoted to Senior Lecturer in 1959, to

Reader in 1962, and to the Chair of Solid State Physics 1966. The new staff that he recruited

(D. Griffiths, A.D. Caplin and B.D. Rainford) broadened the experimental skills that could be

deployed to include electron paramagnetic resonance and neutron scattering. In addition,

Bryan persuaded the department to bring in a succession of young solid-state theorists (S.

Doniach, M. Zuckerman, D. Sherrington (F.R.S. 1994) and N. Rivier), whose achievements

were such that sooner or later they were lured elsewhere. They complemented the theorists

working in the Mathematics Department (now under E.P. Wohlfarth), and, perhaps most

importantly of all, shared office space and coffee room with the experimentalists.

The sense of excitement was heightened by the sabbatical (A.I. Schindler, M. Brodsky, C.

Schiffman and S. von Molnar) and postdoctoral Americans (Z. Fisk, L.L. Hirst and C.L.

Foiles), attracted by Bryan’s outgoing personality and the effervescent mix of experiment and

theory. Many others came for shorter periods, so many that Bryan was once heard to remark

that they may have confused the department with the nearby West London Air Terminal.

The group as a whole continued to maintain a high and distinctive profile in the

international magnetism community into the 1980s, but then changing interests and

departures and arrivals of staff, combined with other pressures on Bryan’s time, encouraged

him to move to a more extramural and collaborative mode of research, much of it

compressed into summer vacations. After his retirement in 1991, and with it the shedding of

administrative tasks, he became Emeritus Professor and a Senior Research Fellow in the

department. He then had the freedom to spend every summer at Los Alamos, and resumed a

vigorous research effort that was maintained right up to his death.

C

Because Bryan had spent his entire academic career at Imperial College and for a time had

been President of the local branch of the Association of University Teachers, he knew the

College structure in all its intricacies. He was therefore a natural choice in 1984 to be elected

Dean of the Royal College of Science within Imperial College, with academic responsibility

for the science departments.

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 8: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

Bryan Randell Coles 57

Bryan became the Deputy Rector of the College from 1986, a post he retained until his

retirement in 1991. There was initially some feeling that with his concentration on research

and teaching he might have been miscast in that role. However, he rapidly gained the

confidence of the whole community. He contributed greatly to the developing strategy of

Imperial College, notably to the decision to expand the commitment to the Life Sciences

initially by merging with St Mary’s Hospital Medical School.

The role of Deputy Rector involved him in the annual round of academic promotions, a

task in which his wide knowledge of the sciences came into full play. His responsibility for

signing research grant applications caused him to take the trouble to read them—sometimes

to the chagrin of the applicants. He also took responsibility for devising a formula for the

allocation of resources to departments. This being a time of severe pressure on budgets,

Bryan’s intellectual and diplomatic skills and sense of humour were called fully into play. He

succeeded in all these tasks because the community had gained total confidence in his

integrity and his humanity.

Humanities

With his own wide range of interests, it was natural that Bryan should seek to encourage the

arts and humanities on campus, but in addition he welcomed the broadening of the College’s

teaching and research base. He had hopes that Imperial College would acquire a similar

standing to that of the Massachusetts Institute of Technology in at least the economic and

social sciences, so he was an enthusiastic supporter of the establishment in 1970 of the

Department of Industrial Sociology, and the recruitment of first Joan Woodward and later

Dorothy Wedderburn.

Similarly, Bryan encouraged the setting up in 1972 of Associated Studies, which before

long became the Department of Humanities, and recruited David Raphael as its academic

director. He helped to nurse the infant department and to overcome the opposition from a

small minority of heads of department to allowing their undergraduates to opt for courses in

literature, history and philosophy.

Bryan’s links with the publishing world made him acutely aware of the need to bridge the

gap between science and the media. He encouraged the College to launch in 1991 the first

MSc in Science Communication to be taught in Europe, preparing science and engineering

graduates for communication careers in the media, and linking closely to the Science

Museum.

P

Bryan’s associations with the scientific publishers Taylor & Francis started when Hume-

Rothery and he wrote their review article (3) for Advances in Physics, which had been founded

and edited by N.F. (later Sir Nevill) Mott, F.R.S. When Mott stepped down from that

editorship he asked Bryan to replace him; the latter’s networking skills in recruiting authors

ensured that the journal continued to flourish.

His involvement with publishing then expanded; he became a member of the board of

directors of Taylor & Francis in 1972, and its chairman in 1976. The explosion of scientific

publishing was soon followed by an era when rapidly increasing production and subscription

costs confronted ever more constrained library budgets, particularly in the UK. Bryan

successfully steered the company through these choppy waters.

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 9: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

58 Biographical Memoirs

This experience and his extensive international contacts made him ideally suited to be

Chairman of the Publications Commission of the International Union of Pure and Applied

Physics (1975–78). He became their representative (1978–82) on the International Council for

Scientific and Technical Information, and from 1983 onward was a member of the latter

organization’s executive board. It was natural that he should become Chairman of the Royal

Society’s Scientific Information Committee from 1993 onwards. He was a keen supporter of

the European Physical Society and a founding member of its Council; he played a major role

in its Publications Commission from 1976 to 1979.

R

Bryan’s half-century-long research career started at a time when, in the aftermath of World

War II, solid-state physics was just receiving recognition as a discipline in its own right. The

foundations for an understanding of the electronic structure of crystals had been firmly

established in the 1930s, in terms of electronic energy bands, Brillouin zones and Fermi

surfaces. However, relevant experimental evidence was sparse, limited for the most part to

measurements of a few macroscopic parameters, such as the electronic heat capacity and the

magnetic susceptibility, which are controlled by the density of states at the Fermi surface, and

to the Hall coefficient, which in the simplest free-electron model is inversely proportional to

the conduction electron density.

Having been schooled in metallurgy, and influenced strongly by Hume-Rothery, Bryan

asked himself, ‘In what way are the physical properties of a metal changed when a certain

amount of another metal is added, and how far can the change in properties be related to a

change in the number and configuration of the electrons present?’ (8). This approach was

therefore very different from that of the low-temperature physicists, who were beginning to

glimpse Fermi surfaces, but initially only in the simple (i.e. non-transition) metals and after

purifying them to the highest degree attainable. There was another factor that separated these

two strands: Bryan was already fascinated by the extraordinary range of behaviour that canbe found in metals and alloys, which is the result of electron–electron and electron–phonon

(and perhaps other) interactions and can often be fine-tuned by variation of the temperature,

magnetic field, pressure, composition and so on. He therefore found little of interest in that

small minority of the metallic elements in which such effects are too weak to tempt them away

from a rather dull non-magnetic and non-superconducting existence.

Bryan felt that chemistry and metallurgy provide plenty of clues as to how electrons in

alloys might behave, in terms of the energies of different electron states and the

configurations of free atoms and ions, and their propensity to alloy with each other or to

form intermetallic compounds of specific structure; there was no point to a first-principles

calculation from an idealized model to demonstrate, say, the complete mutual solubility of

copper and gold. Not all his colleagues had the benefit of his background, as he hinted rather

heavily: ‘Unfortunately, many students of physics nowadays have only limited awareness of

chemistry….’ (14). In his inaugural lecture (8) he felt free to issue the stern injunction to

solid-state physicists: ‘Let none enter who knows not the Periodic Table.’

His strong links with Hume-Rothery continued throughout the latter’s lifetime; Bryan

assisted and later co-authored successive editions of Atomic theory for students of metallurgy

(12); Bryan’s own textbook (with A.D. Caplin) (14) continued the style with updated material.

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 10: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

Bryan Randell Coles 59

When Bryan met Berndt Matthias during an early visit to the USA, the two immediately

found common ground. Matthias’s instinct and insight for the behaviour of transition metals

was a particularly strong influence on Bryan’s approach to research and experimentation.

Bryan’s mathematics was strong enough that he never felt overawed by theory, but he was

not content until he could elicit also a description in terms of a straightforward physical

model. His friendship with N.F. Mott became increasingly important in the second half of his

career. Mott’s own immense stature as a theorist, and his ability to provide deceptively simple

and pictorial explanations, reinforced Bryan’s own suspicions of bull-at-a-gate (or keyboard-

at-a-Cray) theorizing. Mott became Senior Visiting Fellow at Imperial College after his

retirement from the Cavendish Laboratory, and would spend a day or so each week in South

Kensington. Mott’s enthusiasm and eagerness to learn about the latest results, and his catholic

approach to physics, gave an added dimension to his mentoring and friendship with Bryan

and others in the group.

Bryan had a particular affection for resistivity measurements, not just because they could

be done rapidly (although that was a decided advantage), but also because of their sensitivity

to small changes in electronic behaviour as a function of field or temperature. On his arrival

at Imperial, he unearthed from a cupboard a six-dial Diesselhorst potentiometer, a brass,

copper and ebony masterpiece of 1920s electrical instrumentation. It continued to give

sovereign service into the 1980s under Bryan’s curatorial eye; the Tinsley technician who as an

apprentice had helped to construct it emerged annually from his retirement for a service visit.

Bryan learnt the utility of many other experimental techniques, starting with the structural

tools of the metallurgist, and then those for various electrical transport, thermal and

magnetic studies. He was quick to perceive the value of new approaches, such as neutron

scattering and muon spin resonance, if necessary enlisting younger hands and heads, but also

acquiring considerable expertise of his own. Thus he became closely involved with the Institut

Laue-Langevin in Grenoble, and served on its scientific council from 1978 to 1985. At the

same time, he was Chairman of the Science and Engineering Research Council’s Neutron

Beam Research Committee, and had a key role in the establishment of the Spallation Neutron

Source at the Rutherford Laboratory.

Bryan’s strengths in research were his ability to recognize significant conceptual issues and

formulate the questions to probe them, then to identify the best materials and optimum

techniques with which to study them. He would then construct the mental images to

encapsulate the key physical character and implications of the observations, and publicize

them to other experimentalists and to the theorists; in contemporary jargon, he was an

enabler and facilitator par excellence. As communication was also important, the message

would be broadcast with a memorable ‘sound-bite’, drawn perhaps from Bryan’s immense

store of literary quotations, but often original:

Year after year new data come to light,

Effects quite novel make us think again;

Light on old things makes some dark corners bright,

Except that some make older theories vain.

(From Closing Remarks at the International Conference on Strongly Correlated Electron Systems,

Goa, India, 1995 (unpublished).)

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 11: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

60 Biographical Memoirs

Metals and alloys

In the early years of solid-state physics, metallurgical phase diagrams offered important clues

to the understanding of electronic band structure. Hume-Rothery noted that in alloys

between elements that are not too dissimilar in size and electronegativity, the phase

boundaries between different crystal structures occur close to well-defined electron–atom

ratios. Thus the face-centred cubic structure is the stable phase of Cu–Zn, Cu–Al and Cu–Ge

alloys up to an electron–atom ratio of about 1.4:1, and at a ratio of about 1.5:1 the

body-centred structure becomes stable. H. Jones (1934) pointed out that these numbers could

be understood in terms of the relative sizes of a (free-electron-like) spherical Fermi surface

and the relevant Brillouin zones. The inference drawn was that, for these alloys at least, the

electronic band structure is unaltered by alloying, so that they were ‘rigid’ bands (at the time

referred to as ‘collective electron theory’), and that the effect of adding a polyvalent element

to Cu is simply to raise the level of the conduction electron sea, and so inflate the Fermi

surface steadily.

Bryan’s thesis work at Oxford had demonstrated from this kind of phase-diagram analysis

that the rigid-band model gives a poor description of transition metal alloys. In retrospect this

is not too surprising because they are clearly far from free-electron-like, and often have the

added complication of ferromagnetism, but at that time this had not been much understood

by the physicists. These inconsistencies and discrepancies encouraged him to delve into the

theoretical approaches to magnetic alloys then being developed by E.C. Stoner, F.R.S., and

E.P. Wohlfarth.

On arrival at Imperial College, Bryan’s first priority was to see how much more could be

learned about these alloys by complementary studies of physical properties. He still had no

liquid-helium facilities necessary for the measurement of the electronic specific heat, which

would have given direct access to the electronic density of states. Instead, he concentrated

initially on setting up equipment to measure informative parameters just down to

liquid-nitrogen temperatures: resistivity, Hall coefficient and thermopower. In parallel, he

mined the literature for the additional experimental data that he needed, and his early paper

(2) on Ni–Cu and Pd–Ag alloys illustrates these forensic skills. The two pairs of alloys are

isoelectronic and lie in successive rows of the Periodic Table. In both the transition element is

at the end of its series, so that its d-shell should be nearly full; the adjacent noble metal has a

full d-shell and one further s-electron. The rigid-band model therefore predicts that as the

noble metal fraction is increased, the d-band is filled, and thereafter the electron states at the

Fermi level should be purely s-like. The model works well for Pd–Ag, but not for Ni–Cu; the

conclusion drawn was that there is a significant dissimilarity between electron states around

the Ni and Cu atoms, foreshadowing later ideas about the description of impurity states.

Bryan’s success in applying this physical approach nicely complemented Hume-Rothery’s

own crystallographic and structural perspective and consequently, when the latter was asked

to write an extensive review article (3), Bryan contributed the section on physical properties.

The encyclopaedic knowledge gained from this survey was to be retained by his

near-photographic memory and was to provide an invaluable resource for his collaborators

for the decades to come.

Transition-metal alloys: magnetism and superconductivity

An overriding experimental problem in studies of weakly magnetic materials comes from

ever-present contamination with Fe. At worst it can be there in the form of a ferromagnetic

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 12: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

Bryan Randell Coles 61

second phase, but even as an isolated magnetic ion at a concentration of parts per million, at

low temperatures its susceptibility will dominate and obscure the intrinsic weakly para-

magnetic or diamagnetic response of a metal such as Cu or Pd. However, whether or not the

Fe ion is actually magnetic depends on the solvent; at low concentrations it is non-magnetic

in, for example, Al or Nb. This key issue, of the nature of the electronic states of an isolated

impurity atom in a metal solvent, attracted a great deal of attention in the 1960s. It marked

also Bryan’s definitive shift of professional allegiance from metallurgy, in which alloy concen-

trations of less than 0.1% are rarely of interest, to physics, in which an experimentalist’s ideal

dilute alloy would contain just the minimum number of impurity atoms to yield a measurable

effect, and a single atom would suffice for the theorist.

The key to an understanding of these impurities was provided by Friedel’s (1958) concept

of the virtual bound state (VBS), elaborated by Anderson (1961). If the coupling between the

host conduction electron sea and the electron states of the impurity is weak enough, the latter

retain much of their local character, and with it a high density of states (the VBS is the

solid-state analogue of the Ramsauer resonance that was well known in the scattering of

electrons by noble gas atoms). The hybridization depends on the matrix element between

conduction and impurity electron states; if the former is predominantly s-like (as in simple

metals), and the latter d- or f-like (transition metals, and rare earth and actinide elements,

respectively), the difference in symmetry discourages strong hybridization; a low density of

conduction electron states is another favourable factor. A high local density of states increases

the importance of electron–electron correlation on the impurity site, and so favours

occupation of one spin state only; thus when the transition metal, rare earth or actinide

impurity ion is sufficiently weakly hybridized, it is likely to adopt a spin magnetic moment

close to the free ion value. When the competition between localization and hybridization

swings the other way, the moment will tend to disappear. This Friedel–Anderson scheme well

described what was known about dilute solutions of transition metals in the noble metals.

Bryan and his student J.H. Waszink showed, by using as solvents a series of Cu–Zn alloys of

steadily increasing electron density of states, that the magnetic moments could be made to

disappear continuously (9).

Bryan’s international visibility was enhanced greatly by his contributions on transition

metal alloys (6) to the 1963 International Conference on the Science of Superconductivity. To

simplify somewhat, the low-temperature-physics community had focused on understanding

superconductivity in the non-transition metals, such as Al, Sn and Pb, which could be made

very pure and had had their Fermi surfaces mapped. However, these are Type I super-

conductors with low critical fields, and so are almost useless for applications. Bryan

empathized with those of a more metallurgical bent, who were racing ahead with studies of

transition metals and their alloys. These are Type II materials, retaining their superconduc-

tivity up to high magnetic fields, so opening the way to practical high-field magnets. Con-

sequently, there was a powerful driving force (and matching resources) from the major

industrial laboratories—General Electric, Bell, IBM, Westinghouse and others—for their fur-

ther development.

Bryan was intrigued particularly by the correlation between the systematics of the

occurrence of superconductivity and magnetism in the transition metal alloys, and their

response to small amounts of magnetic impurity—usually the inevitable contamination by Fe.

Bardeen et al. (1957) had shown that a high superconducting transition temperature Tc is

favoured by a large electronic density of states N(EF) at the Fermi level, which is contributed

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 13: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

62 Biographical Memoirs

by the part-filled d-shells of the transition metals. Of course, the electron–electron

interactions must not be so strong as to induce magnetism, which is the fate of most of the 3d

transition elements. Of the 4d elements, Nb has a high density of states; with its alloys,

particularly Nb–Ti and Nb3Sn, it became the basis of all high-field superconducting magnets.

A second factor is the survival of the Cooper pairs that form the superconducting condensate.

Ordinary scattering does them no harm, but spin-flip scattering by magnetic impurities is

fatal. In Nb, its large N(EF) brings the benefit that any Fe impurity forms such a broad VBS

that it is non-magnetic and hence harmless. In contrast, the neighbouring element, Mo, has a

somewhat lower N(EF), which of itself decreases the intrinsic Tc by nearly an order of

magnitude, but even worse it allows Fe impurities to maintain a substantial magnetic

moment, which is so destructive of superconductivity that for many years Mo, inevitably

containing some level of Fe impurity, was believed not to be a superconductor. The presence

of a magnetic moment on the Fe also shows up as an anomalous low-temperature resistivity

minimum (4), which was eventually explained by Kondo (1964) in terms of a logarithmic

temperature dependence of the spin-flip scattering.

Local moments and spin fluctuations

The logarithmic increase in resistivity found by Kondo as the temperature is lowered cannot

continue indefinitely. The escape route is that at low enough temperatures, below the Kondo

temperature TK, conduction electrons cloak the impurity with a cloud of electrons of oppo-

site spin, yielding a condensed state that, on average, has no net magnetic moment. However,

the magnetic moment does fluctuate—‘spin fluctuations’—and these fluctuations lead to a

quadratic temperature dependence of the resistivity. The mechanism is essentially electron–

electron scattering, but whereas in a simple metal this scattering rate is proportional to (T/TF)2

(TF being the Fermi temperature), with the fluctuating impurity it becomes proportional to

(T/TK)2 and so can be enhanced by many orders of magnitude.

When the matrix itself is a metal on the verge of ferromagnetism, the spin-fluctuation

effects become even larger. The classic study made with the samples brought by A.I. Schindler

(11) for his London sabbatical exemplifies Bryan’s approach of virtuoso tuning of the

metallurgy to explore the physics: when Ni is added in small amounts to Pd, there is a rapid

increase in the coefficient of the ‘spin-fluctuation’ quadratic term in the temperature

dependence of the resistivity and an enhancement of the static magnetic susceptibility; the

correlation between the two parameters gave conclusive support of the theory (Engelsberg et

al. 1968). Eventually, at ca. 2.0 at.% Ni, the alloy becomes ferromagnetic. Related effects, but

somewhat more complicated in origin, lead in Rh–Fe alloys to a resistivity that increases

threefold from 1 to 50 K (5), a far stronger dependence than in any other material over this

temperature range, and the alloy is used successfully as a practical low-temperature resistance

thermometer.

Intermediate-valence compounds

For the physicist in Bryan, it was a natural and creative extension to move on from the

transition metals to the rare earths and the actinides, with the f-electrons playing the magnetic

and spin-scattering roles of the transition-metal d-electrons. The contrast between the three

series of elements is instructive: the rare earth f-electrons are rather well localized, so that in

an alloy containing rare earth impurities they hybridize only rather weakly with the

conduction electron sea. Well-defined EPR lines are therefore visible, but with g-shifts and

linewidths that are a measure of the residual interaction (7).

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 14: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

Bryan Randell Coles 63

Most of the rare earth elements conform to what would be expected from their atomic (or

ionic) electron configurations; however, at the beginning of the series, Ce has nominally just a

single f-electron (and therefore one that is readily lost). In the broad overview of rare earth

alloys and intermetallic compounds that Bryan had sketched by 1968 (10), the anomalous

behaviour of Ce was already apparent (the element Yb near the end of the rare earth series,

nominally with just a single f-hole, shows similar anomalies). In the actinides, U, and to a

lesser degree Th, have similarly skittish f-electrons.

The idea was already in the air that the electronic f-shell count in Ce could be intermediate

between 0 and 1 electrons, with the preponderant view that this would be a dynamic average

but whose time-scale was uncertain. This ‘intermediate valence’ concept was kept at the back

of Bryan’s mind, from time to time coming to the fore in discussions, until it reappeared in his

experimental programme fifteen or so years later in the context of heavy fermion compounds

(see below).

Spin glasses

Bryan’s name will always be associated with the term ‘spin glass’. In an alloy such as Cu

containing a small amount of Mn, magnetic impurities interact with each other much more

strongly than just through the ordinary magnetic dipole interaction that would occur in an

insulator. The dominant interaction comes about through long-range ripples in the spin

density of the Fermi sea around each impurity (the Rudermann–Kittel–Kasuya–Yosida

interaction) and so is oscillating in sign; the coupling between any pair of impurities cantherefore be ferromagnetic or antiferromagnetic, depending on the spacing between them.

What is the ground state of such a random array of magnetic impurities? Provided that the

impurities are dilute, the ground state will have as many spins up as down, but what will be

the specific configuration of spins? Bryan understood that it would be unlikely to be ordered,

and noticed that the low-temperature heat capacity was large. Characteristically, he picked up

the analogies with the behaviour of amorphous SiO2, whose heat capacity is much larger than

that of the crystalline form (Phillips 1972) and christened the materials ‘spin glasses’.

Bryan contributed to this topic much more than just an accurate and evocative name (18).

The spin glasses studied initially were dilute alloys such as Ag–Mn (13), with good local

moments in a nearly-free-electron-like host. Replacement of the transition metal by a rare

earth, with a more sharply defined moment, leads to further simplification of the idea (15).

Moving in the other direction, as the concentration of magnetic moments increases, the

system becomes less random, and percolation paths linking nearest neighbours become

important; thus, when the Fe concentration in Au exceeds about 15 at.%, ferromagnetism sets

in (16). The situation is slightly different if the solvent is itself a transition metal, particularly

one that is close to magnetism in its own right (17).

The spin glass problem has been enormously influential in providing a model for many

other kinds of random system with competing, and sometimes frustrated, interactions. The

ideas have permeated into fields far removed from magnetic alloys (see, for example, Mezard

et al. 1987), reaching into memory storage and retrieval in neural networks, where spin up and

spin down are translated into neuron firing or quiescence, and the coupling between spins

maps to excitatory or inhibitory interaction between neurons.

Strongly correlated systems: heavy fermions and superconductors

Bryan’s homing instinct in the Periodic Table was always toward the transition metals, the rare

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 15: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

64 Biographical Memoirs

earths and the actinides, where localization, electron correlations and high densities of states

yield a spectrum of magnetic, superconducting and other less predictable behaviours.

The discovery that some intermetallic compounds incorporating these elements have

enormous electronic heat capacities (see, for example, Arko et al. 1972) attracted Bryan’s

attention immediately, and they continued to fascinate him for the next thirty years (22). The

relevant d- or f-shell electrons of the corresponding pure metal form fairly narrow bands

anyway, because of limited direct overlap and weak hybridization. They are narrowed

somewhat further by dilution in an intermetallic compound, but strong electron correlations

within these bands induce a many-body resonance at the Fermi level, which then enhances the

heat capacity, or equivalently the quasi-particle effective mass, by several orders of

magnitude; this led to the term ‘heavy fermions’.

The behaviour of these materials brought together many of the earlier strands of Bryan’s

interests (19). First, the commonest actors on the heavy fermion scene are the rare earth Ce

and the actinide U, previously encountered in the context of intermediate valence compounds.

Secondly, many of the materials can be regarded as concentrated versions of dilute magnetic

alloys. Consequently, when the temperature is lowered and they head towards their ground

state, they are spoilt for choice: spin glass, some other form of magnetic ordering, supercon-

ductivity, charge density waves, and so on.

An alternative to the spin glass outcome uses the electron–electron correlations that build

up the Kondo ground state; these lead to a strongly correlated band or a Kondo lattice. At

non-zero temperatures the magnetic ions cause strong spin-flip scattering, which in Kondo

fashion increases as the temperature decreases. However, because in an intermetallic

compound the ions are regularly arranged, at absolute zero Bloch’s theorem has to prevail,

and the resistivity must vanish. Thus, another signature of heavy-fermion behaviour is a

resistivity that is very strongly temperature dependent at low temperatures—often a quadratic

dependence, as in spin-fluctuation materials.

The structural quality of these complicated materials is always problematic, but Bryan

never forgot his metallurgical skills, and he was delighted to find stacking faults in UPt3

whose occurrence he had long predicted (20). This gave the possibility that the very small

(antiferro)magnetic moment is in fact associated with the defects, rather than being intrinsic.

Bryan took a special pleasure in systems in which small changes in parameters—

composition, temperature or magnetic field—can radically affect the outcome. One of his

latest enthusiasms was for the intermetallic compound URu2Si2, which behaves as a Kondo

alloy at high temperatures, acquires an antiferromagnetic spin density wave on cooling to

18 K, and finally turns superconducting at 1.1 K (21). It has a heat capacity anomaly at the

18 K transition, but one that is far too large in comparison with the small moment associated

with the spin density wave, a mystery that has still not been resolved.

He was always quick to spot the odd man out, and to pick up the vital clue. Two days

before he died, he was discussing with his St Andrews collaborators the next series of muon

spin resonance experiments on ZrV2, a compound that has a magnetic anomaly at 100 K and

then goes superconducting at 8 K. Bryan had suggested that it would not be a conventional

superconductor, and indeed it is not; magnetic impurities cause almost no depression of Tc.

The diversity of Bryan’s interests in alloys and magnetism was celebrated at his retirement

in 1991 by the forty or so contributions to a special issue of Philosophical Magazine (Caplin et

al. 1991). Five years later, the title of the conference held in honour of his seventieth birthday,

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 16: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

Bryan Randell Coles 65

New Frontiers in Magnetism, Heavy Fermions and Superconductivity, could not have been

more apposite to his continuing enthusiasm and proliferation of ideas.

Bryan’s early and sudden death, when he seemed to be in excellent physical health and

intellectually fully productive—he published about twenty papers in the last two years of his

life—deprived the magnetism community of an instantly recognizable voice (McEwen 1998).

His contributions were always pertinent and perceptive, drawing on his vast store of

knowledge (and sometimes his pocketful of notes scribbled on old envelopes). They were

characterized too by an amiable tone, a witty allusion, and occasionally a hint of

self-deprecation. His warmth and charm ensured that he was always seen as a friend as well as

a colleague and collaborator.

Bryan’s research carried forward Hume-Rothery’s banner. It is fitting, too, to record his

admiration for Michael Faraday (8), which perhaps suggests how he himself would wish to be

remembered:

… the French philosophers of the Enlightenment specified as scientific only generalization and

classification, … the extreme involvement with the thusness of things [is characteristic] of the Prussian

metaphysicals. Faraday is perhaps the greatest example of the middle way, proposing general laws and

making acute observations of particular characteristics….

A

It is a pleasure to thank Merivan Coles and Brenda Roddick for the information about Bryan’s early life, and

Sir Eric Ash, F.R.S., Peter Mee, David Raphael and Alan Swanson about his contributions to Imperial College.

On Bryan’s interests in highly correlated systems and spin glasses, Bob Cwyinski, David Edwards, Keith

McEwen, David Sherrington and others have helped me greatly.

The frontispiece photograph was taken in 1995 and is reproduced by courtesy of Mrs M. Coles.

R

Anderson, P.W. 1961 Phys. Rev. 124, 41–53.

Arko, A.J., Brodsky, M.B. & Nellis, W.J. 1972 Phys. Rev. B 5, 4564–4569.

Bardeen, J., Cooper, L.N. & Schrieffer, L.N. 1957 Phys. Rev. 106, 162–173.

Caplin, A.D., Cooper, J., Ford, P. & Loram, L. (eds) 1991 Phil. Mag. B 65, 1113–1464.

Engelsberg, S., Brinkman, W.F. & Doniach, S. 1968 Phys. Rev. Lett. 20, 1040–1044.

Friedel, J. 1958 Nuovo Cim. 7 (Suppl.), 287–311.

Hume-Rothery, W. 1931 The metallic state—electrical properties and theories. Oxford: Clarendon Press.

Jones, H. 1934 Proc. R. Soc. Lond.A 144, 225–234.

Kondo, J. 1964 Prog. Theor. Phys. Japan 32, 37–49.

Mezard, M., Parisi, G. & Virasoro, M.A. 1987 Spin glass theory and beyond. Singapore: World Scientific.

McEwen, K. 1998 J. Magn. Magn. Mater. 177–181, xiii.

Phillips, W.A. 1972 J. Low Temp. Phys. 7, 351–368.

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from

Page 17: Bryan Randell Coles. 9 June 1926 −− 24 February 1997 ...rsbm.royalsocietypublishing.org/content/roybiogmem/45/51...ANBRY RANDELL COLES 9 June 1926 24 February 1997 Elected.R.SF

66 Biographical Memoirs

B

The following publications are those referred to directly in the text. A full bibliography

appears on the accompanying microfiche, numbered as in the second column. A photocopy is

available from the Royal Society Library at cost.

(1) (2) 1951 (With W. Hume-Rothery) The constitution of the Ni–Mn alloys. J. Inst. Met. 80, 85–88.

(2) (3) 1952 Electronic structures and physical properties in the Ni–Cu and Pd–Ag alloys. Proc. Phys.

Soc. B 65, 221–229.

(3) (5) 1954 (With W. Hume-Rothery) The transition metals and their alloys. Adv. Phys. 3, 149–242.

(4) (26) 1963 Low temperature resistivity behaviour of Mo–Fe, Nb–Fe, Pd–Fe alloys. Phil. Mag. 8,

335–337.

(5) (30) 1964 A new type of low temperature resistance anomaly in alloys. Phys. Lett. 8, 243–244.

(6) (32) Electronic structure and superconductivity of transition metals and their alloys. Rev.

Mod. Phys. 36, 139–145.

(7) (38) 1966 (With D. Griffiths) Non s-state resonance in a dilute alloy. Phys. Rev. Lett. 16, 1093–

1094.

(8) (40) 1967 Solid state physics—in particular metals. Bull. Inst. Phys. Phys. Soc. 18, 374–380.

(9) (42) (With J.H. Waszink) Magnetic properties of dilute solutions of transition metals in

copper–zinc alloys. Proc. Phys. Soc. B 92, 731–738.

(10) (43) 1968 (With L.L. Hirst, G. Williams & D. Griffiths) EPR and magnetic susceptibility in noble

metal–rare earth alloys. J. Appl. Phys. 39, 844–866.

(11) (44) (With A.I. Schindler) Low temperature electrical resistivity of Pd and Pd–Ni alloys. J.

Appl. Phys. 39, 956–957.

(12) (45) 1969 (With W. Hume-Rothery) Atomic theory for students of metallurgy, fifth (revised) reprint.

(Monograph and Report Series no. 3.) London: The Institute of Metals.

(13) (62) 1975 (With W.H. Pon) Magnetic interactions in dilute AgMn alloys. Phys. Rev. Lett. 35, 1655–

1658.

(14) (64) 1976 (With A.D. Caplin) Electronic structures of solids. London: Edward Arnold.

(15) (67) 1977 (With M.H. Bennett) La(Gd)Al2—a simple spin glass? Physica B/C 86–88, 844–845.

(16) (74) 1978 (With B.V.B. Sarkissian & R.H. Taylor) The role of finite magnetic clusters in Au–Fe

alloys near the percolation concentration. Phil. Mag. B 37, 489–498.

(17) (80) 1980 The experimental status of the Stoner spin glass. J. Magn. Magn. Mater. 15–18, 157–

158.

(18) (98) 1985 An historical introduction on spin glasses. Annales Phys. 10, 63–67.

(19) (105) 1987 Heavy-fermion intermetallic compounds. Contemp. Phys. 28, 143–157.

(20) (131) 1993 (With B.G. Demczyk, M.C. Aronson & J.L. Smith) Observation of a secondary structu-

ral phase in the heavy-fermion superconductor UPt3. Phil. Mag. Lett. 67, 85–88.

(21) (142) 1995 (With J.-G. Park) Effects of rare-earth substitutions for U in URu2Si2. Physica B 206–

207, 418–420.

(22) (144) 1996 3d heavy fermions: do they exist? Physica B 223–224, 260–261.

on May 20, 2018http://rsbm.royalsocietypublishing.org/Downloaded from