The Man in the Monkeynut Coat: William Astbury and the Forgotten Road to the Double-Helix

  • Kersten T. Hall
Oxford University Press: 2014. 9780198704591 | ISBN: 978-0-1987-0459-1

History is written by the victors, in science as in war. The story of the race to uncover the structure of DNA was thrust into the public imagination by James Watson's highly partial 1968 account, The Double Helix. Ever since, historians have been filling in the gaps. Most notably, successive authors have corrected Watson's unkind portrait of Rosalind Franklin, whose X-ray-diffraction photograph of DNA ('Photo 51') was crucial to the model developed by Watson with Francis Crick.

Now science historian Kersten Hall has disinterred another figure from the footnotes of the DNA story, and brought him into the limelight. William Astbury — known to all as Bill — was part of the group of X-ray crystallographers who learned their craft at the feet of William Henry Bragg, the father of the field, at London's Royal Institution in the 1920s. Astbury, along with fellow student J. D. Bernal, was fired with the new technique's possibilities for studying the key molecules of life, especially the proteins that are fundamental to all living tissue.

William Astbury used X-ray crystallography to unravel the changeable structures of proteins. Credit: Leeds Univ. Lib.

Bernal left the Royal Institution to found a lab at the University of Cambridge, UK; Astbury was banished (as he saw it at the time) to the University of Leeds in northern England. There he applied X-ray analysis to the study of natural fibres such as wool and hair: Leeds was an important centre of the textile industry. The wool protein keratin, Astbury complained on his appointment, was “lifeless and uninteresting”. Later, he would tell public audiences that wool had given him “a glimpse of the loom on which the web of life is woven”.

In the early 1930s, Astbury obtained X-ray-diffraction photographs of stretched and unstretched wool fibres, which suggested that the protein chain could switch between a compact and an extended form. Apart from shedding light on the elasticity of wool, his discovery showed that the properties of long-chain proteins depended on their three-dimensional shape, and that changes in shape could have functional consequences.

Hall makes a good case for Astbury's contribution to protein-structure analysis, but the book's subtitle signals that the author is really interested in DNA. In 1938, Astbury and his PhD student Florence Bell — a protégée of Bragg's son Lawrence, and now largely forgotten — published the first X-ray photographs of DNA fibres. They were blurred and hard to analyse, because the Leeds team did not know that two forms of DNA were mixed in their sample. But they contributed to Watson and Crick's thinking, and encouraged Maurice Wilkins — and subsequently Franklin — at King's College London to undertake X-ray studies of the molecule.

Astbury's student produced X-ray photos of DNA that were almost identical to Rosalind Franklin's, taken the following year.

In 1951, Astbury's “lab boy” turned doctoral student, Elwyn Beighton, produced X-ray photos of DNA that were almost identical to Franklin's Photo 51, taken the following year. Astbury never published Beighton's pictures or presented them in public, nor did he seem to recognize that the distinctive X-shaped diffraction pattern showed that DNA was a helix.

Hall is clearly frustrated that Astbury missed the chance to win a Nobel Prize and change the face of biology. He dwells at length on what might account for this failure. Astbury had become depressed and embittered as he saw others in London and Cambridge win funding that he had failed to secure. He had conceded the DNA problem to Wilkins at King's. He may not have kept up with the theory of X-ray diffraction. Finally, Hall suggests, Astbury probably saw DNA's apparently fixed structure as uninteresting compared to the conformational changes in proteins; at the time, his main interest (and the subject of Beighton's thesis) was bacterial flagella.

Hall indulges in some counterfactual speculation. What if Astbury had shown Beighton's pictures to the US chemist — and later Nobel laureate — Linus Pauling, who stayed with Astbury in Leeds in 1952? Pauling might have seen their significance and realized Watson and Crick's worst fears by beating them to the structure. But it didn't happen, and Watson's 1968 book credits Astbury only with the “one half-good photograph” taken by Bell in 1938.

Hall tells his story with style and pace. But I am unconvinced by his central contention, that Astbury was a “titan” and the founder of molecular biology. Astbury was one of many who independently promoted a molecular approach to understanding living things. Success in science depends as much on personality as on intellect, and here Hall leaves us largely in the dark. Clearly Astbury was a larger-than-life figure, a good communicator and entertaining in company (if you liked “off-colour jokes”), yet not favoured by the scientific establishment. Crystallography pioneer Dorothy Hodgkin (who taught chemistry to Astbury's daughter Maureen at the University of Oxford, UK — a connection not mentioned by Hall) once described him affectionately as “very bad and very amusing”.

And the monkeynut coat? Astbury worked with the now defunct UK company Imperial Chemical Industries to develop fibres made from a protein derived from peanuts, and proudly wore a coat made from 'Ardil'. It proved to be no cheaper and considerably less hard-wearing than wool, and never took off. Hall uses this story to launch a final chapter exploring Astbury's prescient reflections on the manipulation of biological materials for utilitarian ends. He did not live to see the age of biotechnology, but he was certainly one of its prophets.