The bridge to genius

One of the more fascinating mysteries of the century that has recently ceased to be ours is the extraordinary efflorescence of…

One of the more fascinating mysteries of the century that has recently ceased to be ours is the extraordinary efflorescence of creativity in art and science that occurred in or about the middle of its first decade. The few years after 1905 saw the coming to full power of two of the greatest creative minds of the past 100 years, or indeed, of the past two millennia. In Paris, the young Pablo Picasso produced Les Demoiselles d'Avignon, the picture that was to change the art of painting forever, while in Bern, the Patent Office clerk Albert Einstein formulated the Theory of Relativity, which would overthrow classical physics and usher in the atomic age.

That two such revolutionary triumphs should be achieved in such a short space of time seems more than coincidental. Of course, Picasso and Einstein did not spring fully armed from their own brows. They had their precursors, as well as contemporaries whose endeavours in the fields of both of them propelled them forward toward the resolution of the tasks they had set themselves; as Newton had observed, he would not have been able to see so far had he not been standing on the shoulders of giants.

In this learned, inspired, and daring book, Arthur I. Miller poses a hypothesis to account, in part at least, for the twin phenomenon that was Einstein/Picasso. In general, and wholly persuasively, he shows how both men came to the realisation that art and cosmological physics must be refounded on the basis of geometry. This may seem a more surprising conclusion in the case of Picasso, but for Einstein too it required extraordinary insight to see that, as Miller puts it, the special theory of relativity, which Einstein formulated in 1905, "could be fully generalized only through geometrization".

In particular, and more controversially, Miller suggests that the link between the discoveries of the artist and the scientist was a single man, the French mathematician and theorist, Henri PoincarΘ, and, even more particularly, one work of his, the bestselling book, La Science et l'hypothΦse. PoincarΘ had one of the finest minds in the history of mathematics; Miller, who in Paris in the 1970s discovered the great man's lost papers, points out that PoincarΘ "made significant contributions to all branches of physics and astronomy", that space vehicles are still launched and their orbits calculated according to his version of Newtonian mechanics, and that his work helped lay the foundations of chaos theory.

READ MORE

Pablo Picasso had arrived in Paris from Barcelona in 1904, the penniless son of an academic painter who early on had despaired of his own work when he recognised his teenage son's astonishing talent. The young Picasso, ambitious, self-assured, and ever on the look-out for advances in art that he might assimilate - Miller notes that both Picasso and Einstein "were opportunists, willing to take advantage of every hint offered by the intellectual currents in which they swam" - settled into the famous "Bateau Lavoir", the squalid artists' rooming house on the hill of Montmartre where the hideous Montmartre tower stands today. At once he set about equipping himself with the two essential necessities of the vie de boheme: a mistress, and a dealer. The girl, the opulent, statuesque Fernande Olivier, who was to be the painter's muse, and model, for many years, was easy to find for a young man of Picasso's passion and magnetism. He had to wait some years before the dealer, the canny but immensely supportive Daniel-Henry Kahnweiler, came along.

Picasso was a profoundly serious and dedicated artist, for all his showmanship and self-dramatising gestures - he had acquired the playwright Alfred Jarry's pistol, with which he would shoot, with blanks, anyone whose opinions he disagreed with - and he set to work at once in his chaotic studio in the Bateau Lavoir to create an artistic revolution, and make a work of art that, in Miller's words, "would measure up to the magnificent achievements of science". Taking all he could from such diverse influences as CΘzanne's passionately cerebral experimentalism, the time-lapse photographs of Eadward Muybridge, and an exhibition of African primitive art at the TrocadΘro, he pushed the boundaries of pictorial space beyond all previous limits. He spent years working and reworking the picture that would become Les Demoiselles; when he had finished it, there was hardly a soul who liked it, and some of his best friends laughed at it.

There were some, however, who saw the direction in which he was headed. One of them was Maurice Princet, whom a sarcastic critic later dubbed "the inventor of Cubism". He was an insurance actuary and amateur mathematician who had found his way into la bande α Picasso, the group of the painter's friends and admirers who gathered each evening in the bars and bistros of Montmartre to discuss the latest developments in art and, thanks largely to Princet, in mathematics. It was Princet who introduced Picasso to non-Euclidean geometry, and the work of PoincarΘ; in particular PoincarΘ's theories on the so-called "fourth dimension". Miller quotes a highly suggestive passage from La Science et l'hypothΦse, in which it is hard not to see a prescription for the procedures of cubism:

The images of external objects are painted on the retina, which is a plane of two dimensions; these are perspectives. But as the eye and objects are movable, we see in succession different perspectives of the same body taken from different points of view . . . Well, in the same way that we draw the perspective of a three-dimensional figure on a canvas of three (or two) dimensions, so we can draw that of a four-dimensional figure from several different points of view.

Einstein, too, in Bern, had read PoincarΘ, in German translation, in the course of the proceedings of the so-called Olympia Academy - it consisted of himself and two friends - and had found his theories of interest and use. He was impressed with the Frenchman's work in the area of electromagnetic theory, and, especially, in the area of temporal relativity. Miller quotes a passage from La Science et l'hypothΦse which is as suggestive in the consideration of Einstein's work as the passage above on the fourth dimension was for the development of cubism: "There is no absolute time. To say that two durations are equal is an assertion which has by itself no meaning at all and which can only acquire one by convention." It might be Einstein himself speaking.

For all that PoincarΘ, among others, was an inspiration for him, Einstein was always his own man. Working eight hours a day in the Patent Office, coming home in the evenings to deal with his discontented wife Mileva and a child he had probably not wanted - he had forced Mileva to abandon their firstborn, a daughter, whom he had never even seen; in similarly cruel fashion, Picasso and Fernande Olivier at one point adopted a daughter, but quickly tired of her and handed her back to the orphanage - he still managed to produce, at an extraordinary pace, a series of papers of unparalleled beauty, simplicity, and rigour, instituting in the process an entirely new conception of physical reality.

What is perhaps most remarkable about Einstein the scientist - as a man, he was less than a paragon - is his ability to think in entirely new ways about problems that had been baffling scientists for generations. He approached these problems with an almost childlike curiosity, and a playful delight both in the challenge and in his own powers to meet it. In later life he recalled, with a sort of wonderment, how one of his fundamental early discoveries came to him.

"I was sitting in my chair in the Patent Office in Bern when all of a sudden a thought occurred to me: 'If a person falls freely he will not feel his own weight.' I was startled. This simple thought experiment made a deep impression on me. It led me toward a theory of gravitation." Here is Miller:

He thought about these issues in a manner strikingly different from other physicists. He based his ideas on concepts from our daily movement about the world, he later noted, because "scientific thought is a development of pre-scientific thought" . . . Einstein wrote, "the whole of science is nothing more than a refinement of everyday thinking. [The scientist] cannot proceed without considering critically a much more difficult problem, the problem of analyzing the nature of everyday thinking.

The "genius" Albert Einstein and l'homme moyen sensuel Leopold Bloom were not so very different after all.

Einstein, Picasso is a fine and stimulating book. It makes breath-taking connections across the gap between the so-called two cultures of science and art, identifies the mysteriously synchronous effects of the Zeitgeist, places aesthetics firmly at the heart not only of art but of science as well, and, not least, celebrates the excitement and the glories available, even yet, to the life of the mind. It is also, and not least, a timely and much-deserved tribute to the great Henri PoincarΘ. As to whether or not Miller has proved his central premise, that PoincarΘ was a vital influence on both Picasso and Einstein, the reader will have to make up his or her own mind. Probably the finest sections of the book are those which detail Picasso's titanic struggle to make the breakthrough that is Les Demoiselles, but for anyone who has not so far ventured into the heady and refreshingly non-commonsensical world of relativity theory, here is the place to start.

John Banville is Associate Literary Editor and Chief Literary Critic of The Irish Times