You have /5 articles left.
Sign up for a free account or log in.

Now that “genius” has become the job title for the person who fixes your MacBook, we need something considerably stronger to describe the Indian mathematician Srinivasa Ramanujan. Awe seems like the only suitable response to the work Ramanujan did and how he did it.

He was born in the southern part of the country in 1887, one year following publication of A Synopsis of Elementary Results in Pure Mathematics by George Shoobridge Carr, a math tutor in London. The volume would have been long since completely forgotten had Ramanujan not come across it as a high school student. Carr assembled more than 6,000 formulas and theorems in order of growing complexity -- but without the full proofs. Those Ramanujan worked out for himself.

By his twenties, Ramanujan was filling notebooks with his own extremely advanced work in pure mathematics, samples of which he sent to G. H. Hardy, an eminent number theorist at Cambridge University, in 1913. Following the example of Carr’s Synopsis, Ramanujan presented his findings without spelling out the proofs. He also used notation that had grown out of date, and it is easy to imagine the Cambridge don throwing the letter with its attachments into a drawer, along with all the other pleas for attention from amateur mathematicians. Instead, Hardy examined Ramanujan's material, found it interesting and in some cases staggeringly original, and helped wrangle the fellowship that brought the young Indian savant to Cambridge in 1914.

Ramanujan spent the most of the remainder of his short life in England, immersed in finding or inventing whole new domains of mathematics, even as tuberculosis undermined his health. Whether mathematicians discover concepts (as astronomers do galaxies) or create them (as composers do symphonies) is a matter of perennial controversy; for his part, Ramanujan said that ideas came to him in dreams sent by the Hindu goddess Namagiri. However one understands that claim, much of the work was so advanced that his colleagues were barely beginning to catch up when he died in India in 1920, at the age of 32.

The effort continues. Ken Ono's My Search for Ramanujan: How I Learned to Count (Springer) is the memoir of a mathematician who has devoted much of his career to working out the proofs and methods that his predecessor left unstated. And the story would be interesting enough as such, even if the author's life did not have its own twists and turns. Ono, a professor of mathematics and computer science at Emory University, wrote the book in collaboration with the late Amir D. Aczel, best known as the author of Fermat's Last Theorem. The input of a capable historian and popularizer of mathematics undoubtedly helped Ono create a smooth a compelling narrative out of extremely difficult material.

By anyone else's standard, Ono was, like his siblings, a gifted child, although fate seems to have rendered his talents a burden. His parents emigrated from Japan in the 1950s, and the author recalls his own childhood in the 1970s as defined by a "confusing and frustrating intersection of incompatible cultures." Even harder to reckon with was the unmeetable standard of Olympian intellect embodied by his father, Takashi Ono, a professor of mathematics (now emeritus) at Johns Hopkins University. As for his mother, Ken Ono describes her as "present[ing] herself as a martyr who had sacrificed all self-interest for the family," thus "instilling in us a sense of duty to succeed in the lives that they had planned for us."

And planned with unforgiving precision, it seems: his parents' only friends "were other professors with overachieving children who were being accepted by top private colleges and winning elite music competitions," establishing "models of perfection" that Ono and his brothers were reminded of constantly. He describes his parents as carrying the tiger mom outlook (that "if their children are not at the top of their class, then the parents aren't doing their job") to such an extreme that not even academic achievement merited praise. While anything less than perfection brought shame upon the family, mere excellence hardly merited notice.

One brother is a now a biochemist and university president, the other is a music professor, and Ken himself has an imposingly long list of professional achievements. Judged simply by the results, then, the Ono parenting style was a success. But the cost was enormous: decades of anxiety, self-doubt and self-contempt, taking him to the verge of suicide. The sight of math prodigies so young that their legs didn't touch the floor when they sat down in the classroom made passing advanced undergraduate courses feel like proof of inadequacy. Harsh and unrelenting parental voices echoed in his head ("Ken-chan, you no can hide …. You must be one of the best, and right now you losing out to 10-year-old kid with Pac-Man watch").

But the push to overachieve also met inner resistance. He engaged in competitive bicycling and played gigs as a disc jockey, and it sounds like there were enough fraternity shenanigans to feel liberated from what Ono calls "my old image as Asian-American math nerd." He had brushes with what would count as academic humiliation even by standards far less exacting than his own. But behaving "like a goofball" (in the author's preferred expression) seems, on the whole, to have been therapeutic. Ono eventually received his Ph.D. -- an achievement his parents took as a given and so never commented on.

One remarkable thing about Ono's narrative is that he seldom, if ever, sounds angry. To understand is to forgive, the proverb runs -- and coming to an intellectual comprehension of one's parents' outlook and behavior is a necessary step toward dealing with the consequences. (The second- and third-generation offspring of immigrants often have to come to terms with how the first generation navigated the unfamiliar or hostile circumstances they faced.) But in Ken Ono's case, there is another, equally compelling force: a series of encounters with the example and legacy of Ramanujan -- sometimes accidental and, at other times, sounding very much like destiny. I am reluctant to say much more than that because the part of the book's emotional power comes from the element of surprise at how developments unfold. Suffice it to say that mathematics, which for obvious reasons Ono came to consider an unpleasant and compulsory part of his lot in life, comes alive for him with all the beauty and mind-blowing glory that Ramanujan implied in referring to the goddess.

But that revelation has a much more human aspect in Ono's memoir, which is an account of the life-enhancing (and quite possibly life-saving) influence of a few friends and mentors. When G. H. Hardy responded to Ramanujan's letter in 1913 and fostered the promise of his early work, it saved a genius from the threat of oblivion and made possible an extraordinary flourishing of mathematical creativity. It will not give too much away to say that My Search for Ramanujan tells a comparable story, and does so in a way that pays tribute to collegiality as something more than a form of courtesy.

Next Story

Written By

More from Intellectual Affairs