The Boy Who Played with Fusion

Home > Other > The Boy Who Played with Fusion > Page 11
The Boy Who Played with Fusion Page 11

by Tom Clynes


  In athletic pursuits, certain “early specialist sports”—gymnastics, figure skating, diving—can produce thirteen-year-old Olympic standouts, but star turns in team sports typically come later. Among intellectual pursuits, writing contrasts with mathematics and chess in that signs of literary mastery appear much later. “Writers need more human experience to understand and incorporate into their craft,” says David Henry Feldman.

  An early interest in science is one of the major predictors of eventual success in scientific careers, but the cutoff points for acquiring interest and skills in the sciences are less predictable than in other fields, and Kerr’s research has shown that gender expectations can thwart the development of scientific talent. “When a child shows a passion for something that a parent fears isn’t consistent with other children of their gender,” she says, “some parents still fear that their kids will be bullied for being ‘different.’” Instead of supporting a child’s passions, a parent might push team sports on a boy who’s more interested in theater or art, or shower princess dolls on a girl who likes science or athletics.

  The impact of those parental choices are amplified in the lives of gifted children, Kerr says, “because the double messages are so strong in terms of achievement versus preferred sex role. I counsel parents to err on the side of achievement, but it’s very difficult. Parents remember their own childhood experiences, and how severely kids punish other kids for not being obedient to gender roles.”

  A parent is more likely to notice and feed childhood interests if they’re related to the parent’s own career. Subotnik surveyed children at specialized science high schools and found that 60 percent had one or both parents in a scientific field. It’s harder if you’re a child with a talent in an area that your family doesn’t care about. “In immigrant families especially, children sometimes get negative reinforcement if they don’t want to be, say, a doctor or an engineer,” says Subotnik.

  Psychologist Scott Barry Kaufman, a researcher at the University of Pennsylvania, warns against closing the door too early on the “seemingly untalented” who are not early bloomers, and he uses his own experience as a compelling example. Diagnosed with an auditory processing disorder, he was sent to a school for children with learning disabilities and, he says, “fed a steady stream of low expectations.” When he was fourteen, a new teacher noticed his boredom and frustration and asked why he was still in special education.

  “For the first time in my life,” Kaufman says, “my mind was suddenly brimming with possibility as I wondered: what am I actually capable of achieving?” He now has a master’s in philosophy from Cambridge and a PhD from Yale and is the author of the book Ungifted: Intelligence Redefined. Kaufman readily admits that his story is a single anecdote and that he may be “just an outlier.” But because a child’s gifts may not be readily observable at any single moment in time, Kaufman says it’s “an egregious error” to label and sort children as “gifted or ungifted” too early or to suggest limits on what a child can ultimately achieve.

  The trick, says Kaufman, is to find a creative outlet that best fits a child’s unique set of characteristics. “Once you find that domain, the learning process can proceed extremely rapidly, as the individual becomes inspired to excel.” He recommends that parents stay on the lookout for activities that bring on the state of focused motivation that Hungarian psychologist Mihály Csíkszentmihályi famously termed flow—a single-minded immersion in an activity in which one becomes so motivated and involved that nothing else seems to matter.

  In Taylor’s case, his enthusiasm came so early—and was so focused, so intense, and often so vocal—that it was impossible to miss. Working in his laboratory, he’d regularly enter the flow state, becoming so absorbed in his experiments that he’d forget about time, food, and everything else. It was relatively easy at the outset for his parents to nurture their son’s evolving obsessions, even when he first got into nukes. But what had started with their son’s science-fair assemblage of dinner plates and lantern mantles had evolved into his desire to build an atom smasher with the power to throw out several varieties of DNA-mangling radiation. Increasingly, Tiffany and Kenneth seemed to be faced with a choice between supporting their son’s passions and keeping him alive. At some point, they realized, they might have to draw the line.

  For teenage rocketeer Homer Hickam’s father, that point came after his son’s errant rocket roared through town, narrowly missing several miners and slamming into the mine’s headquarters. The boy’s obsession with rocket science had been inspired by the Sputnik launch and Wernher von Braun’s response. Homer and his friends had captured the townspeople’s imagination and moral support as their rockets grew bigger and went ever higher. But they had also raised hell with a litany of botched launches, busted fences, fires, and explosions. To Homer’s father, the mine’s foreman, his son’s hobby represented an embarrassing and dangerous lack of discipline.

  The inadvertent rocket attack on the headquarters convinced the elder Hickam that it was just too dangerous—for the miners and their families, for his son, for his job and reputation—to let the launches continue. That evening, he came home, gathered Homer’s chemicals into a box, and took it and his son to a nearby creek.

  “This is the end of it,” he said, pouring everything out. “And this time I mean it. Collect stamps, catch frogs, keep bugs in a jar, do whatever you want. But no more rockets.”

  That Sunday in church, the reverend looked directly at Homer, who slid lower in his pew, and quoted from the Bible: “A foolish son is the calamity of his father. Cease, my son, to hear the instruction that causeth to err from the words of knowledge.”

  What Homer didn’t realize yet was that the teachers and women’s club had lobbied the reverend to do whatever he could to keep the town’s young scientists in business. The preacher again quoted the Bible: “He that begetteth a wise child shall have joy of him. To have a child who longs to learn is the sweetest gift of all.”

  The reverend, on a roll, continued:

  Sons, obey your fathers. But fathers, help your sons to dream. If they are confused, counsel them. If they stray, search them out and bring them home . . . Fathers, I beseech you to seek out your straying sons and rescue them by keeping their dreams alive. These boys, and we all know I’m talking about our very own rocket boys, are dreaming great dreams. They should be helped, not stifled.

  The sermon turned the momentum back toward the rocket boys. After church, the elder Hickam tossed his son the car keys and directed him to a slack dump outside of town. Tons of coal tailings had been bulldozed to create a square mile of flat black desert, perfect for launching and recovering rockets and far enough from town to be safe.

  Cape Coalwood would serve as the proving ground for the club’s increasingly ambitious science-based rocketry program. Their ingenious designs would go on to win the National Science Fair, and Homer Hickam, the boy rocket scientist, would go on to Rocket City (Huntsville, Alabama) to become a NASA aerospace engineer.

  Kenneth and Taylor arrived at the Coke plant and walked through the cavernous space that was morphing, week by week, from a bottling factory to a warehouse. They ran into Tom Chesshir, the service manager who repaired the company’s vending equipment. Chesshir, in his sixties at the time, was a big man, an outgoing ex-athlete and coach who, as Kenneth puts it, “never met a stranger.” The nature of his job—half electrician and half mechanic—and his farming background made him a great all-around handyman.

  When Taylor mentioned his would-be project, Chesshir remembered that he’d once seen a Van de Graaff generator in operation at a science fair in Dallas, where he’d taught earth sciences for five years. “A guy put his hands on it when it was running, and his hair stood up on end,” Chesshir said. “Just about the strangest thing I ever did see.”

  “And it didn’t kill him?” Kenneth asked.

  Chesshir explained that even though a Van de Graaff machine can generate hundreds of thousands of volts, the amperage
(the measure of electrical current) is too low to be hazardous.

  “Current is what kills,” Chesshir told Kenneth, “not voltage.”

  It was something that Kenneth had missed or that Taylor had neglected to mention in his hyperkinetic sales pitches to his father, whose head was spinning with all the talk of kilovolts and gamma rays and neutrons. Taylor unrolled his blueprints of the Van de Graaff generator for Chesshir and took him through the design—leaving rolled up, for now, the plans for the other half of the project, the accelerator that physics pioneers had used to smash subatomic particles and release some seventeen million electron volts.

  Chesshir was one of Kenneth’s most trusted employees. “What would you think,” Kenneth asked him as they looked over Taylor’s plans, “about helping Taylor build this thing this summer?”

  Chesshir liked Taylor and enjoyed working with young people. He said he’d love to help. He and Kenneth made a deal that Chesshir would come in a day or two each week, as his other duties allowed, and work with Taylor in a corner of the plant, using as much scrap equipment as they could.

  Taylor, ecstatic, ran over to a pile of machinery he’d collected from the bottling line and showed Chesshir and Kenneth a few of the salvaged O-rings, knobs, and other equipment he thought would be suitable for the project. He pulled out an instrument panel with switches on the front and an octopus tangle of wires tumbling out of the back.

  “And this,” Taylor said, holding it up, “will be our control panel!”

  12

  * * *

  Heavy Water

  TAYLOR, WHO HAULED IN PARTS and tacked his diagrams to a wall in the corner of the plant’s workshop area, says his role was lead designer. Chesshir’s role, which he’d discussed privately with Kenneth, was to help Taylor find parts and build the machine and, mostly, Chesshir remembers, “keep Taylor from blowing us all up.”

  Three-quarters of a century before Van de Graaff invented his machine, James Clerk Maxwell published his landmark equations on electromagnetism. Until then, scientists had wondered how an electric charge could exert forces on distant charges and objects, how such a force moved across empty space, and how fast it traveled. Maxwell’s explanations made it clear that electricity and magnetism were best understood as fields, which fill surrounding space and indirectly affect other charges and objects within the field.

  A conventional Van de Graaff machine creates an extreme electrostatic imbalance by using a sphere atop an insulating column that separates two electrodes that are charged by a motor-driven belt. The belt pulls electrons from one terminal and dumps them in the other, building up a strong electric field between the terminals.

  Taylor was just getting comfortable using power tools, so Chesshir guided him as they cut large pieces of metal and wired everything together. Taylor had hoped to build his machine using mostly scrap materials from the bottling plant. But once they got started, he and Chesshir ended up making trips to the hardware store or ordering something online nearly every day. They picked up a pair of extra-large stainless-steel mixing bowls at Walmart that became the hemispheres for the globes after they cut the lips off and joined them with aluminum foil tape. The trickiest part was the belt, which had to be custom made from neoprene strips.

  “Taylor was always go, go, go,” Chesshir says, “never wanting to stop and take a break. It was fun watching someone that ambitious with that kind of attention span, which is unusual for a kid today. But he wanted every step of the project done that day, right then. I’d have to tell him, ‘Look, we’re in Nashville, Arkansas, five thousand population; we’re just going to have to send away for this.’ That wait for a delivery would just about kill him.

  “When we finally got the generator going,” says Chesshir, “that was a big day! Here was this kid I’d known since he was two, when he was too shy to talk, and suddenly when that thing started sparking, his eyes were lighting up and he was jumping for joy.”

  One way to estimate voltages is by measuring by how far a spark will jump through the air. A static charge of ten thousand to thirty thousand volts jumps a half an inch or so; Taylor’s machine could throw a spark five or six inches. “We had a whole lot of fun testing it out by shocking each other,” Taylor says. “You’d get close and you could feel your arm hair standing up. I’d have Tom or Joey put their hands on it, then I’d run up and shock them, and they’d do the same to me. Even though the charge is small, it’s enough to make you jump and give a little yelp.”

  With the power source finished, they turned to building the particle accelerator; here, Chesshir reached the limits of his expertise. “When I’d try to get Taylor to explain it,” Chesshir says, “he’d give a big dissertation about subatomic particles and neutrons and I’d say, Never mind, just tell me what we need to get.”

  Taylor had managed to track down Larry Cress, the source of the plans for the proton/deuteron accelerator featured in the 1971 Amateur Scientist column. After they talked on the phone, Cress sent Taylor a big packet of journal articles and design documents.

  Taking cues from the trove of materials, Taylor modified the plans. Lacking a TIG (tungsten inert gas) welder to connect aluminum and other metal parts, Taylor and Chesshir built the accelerator column out of glass, copper, and brass and attached it to a collecting dome identical to the one atop the Van de Graaff generator.

  The original plans specified an ignition coil from a Model T Ford to generate the arc for the ion source. In this case, Taylor was able to stay true to the plan. “My grandfather had his Model T parked in the warehouse, and I asked him if I could borrow the spark coil from it. It turned out to be perfect for the job. It was high voltage and compact, so I could easily fit it inside the small space inside the accelerator.”

  If all went according to design, the arc would strip the electrons off atoms, creating plasma, the ionized gas consisting of free electrons and positive ions. These ions would then accelerate down the tube and collide and produce gamma rays (if they used a hydrogen source) or neutrons (if they used deuterium).

  As the project took shape, Chesshir began to trust Taylor’s judgment more and more. “I was impressed by how he was thinking everything through, improving the original plans. But some of that stuff I was ordering online—well, let’s just say Homeland Security’s not doing their job. I sometimes worried I might get arrested, because no one would believe I was just the gopher and it was an eleven-year-old running the show.”

  To produce neutrons, Taylor needed deuterium gas to fuel the accelerator. Deuterium, a stable isotope of hydrogen, is not radioactive. “But my parents equated hydrogen with the Hindenburg disaster,” he says, “and they weren’t confident enough to buy me a canister of deuterium gas. So I had to figure out a way to make my own.”

  Taylor ordered some heavy water (also called deuterated water) from United Nuclear Scientific Supplies. Heavy water, made from oxygen bonded to one or more deuterium atoms, is used in Canadian nuclear reactors to moderate the unenriched uranium fuel. To extract deuterium from the heavy water, Taylor modified a Vacutainer blood-collection tube and turned it into an electrolysis device that could split the liquid into deuterium and oxygen. “It was a pretty sophisticated little rig,” he says, “and it worked really well.”

  The finished accelerator looked great. “When I compare it to pictures of the first cyclotron that Ernest Lawrence built [for which he won the Nobel Prize in 1939], I can see that our construction techniques are similar—although I think mine looked even better. We had that nice control panel off the bottling line, with lots of cool-looking buttons and switches and relays,” Taylor says. “We had it wired up so it would start the pumps and the ion sources.”

  But Taylor hadn’t realized how challenging it would be to produce a workable vacuum. For his particle accelerator to function correctly, he’d need to pump out most of the air inside the accelerator column to create an almost empty space for his subatomic particles to travel in. If any gas or air molecules were left inside the tube, th
e accelerated particles would collide with them and lose energy. “Imagine a freeway in Los Angeles and you want to go a hundred miles an hour,” Taylor explains. “If you try that at rush hour, you’re going to hit other cars. But in the middle of the night, it’s wide open and you can go fast.”

  To pump the air out of the tube, Taylor used a refrigerator compressor and wired it to run backward. Then he opened a valve to inject a small amount of his homemade deuterium gas. “I was so excited,” Taylor says. “Me and Tom got the generator up above two hundred thousand volts, and with the Model T arc, we tried to get plasma going.” But even though they used higher-tech fasteners than Lawrence had in the 1930s, they couldn’t create a good enough vacuum to sustain a plasma field or clear enough air molecules to accelerate particles to any measurable degree. They tweaked the fasteners and tried all sorts of sealants—silicon rubber, epoxy, “and a few other things,” says Taylor. “We were using techniques from the sixties and seventies, and we modernized them, but with our expertise and the materials we had, we could only go so far. Most of it worked. But not the big picture.”

  Taylor didn’t know it then, but the setback was probably among the best things that could have happened to him in the long term. By taking on a challenge that was beyond him, he was, likely unconsciously, putting the value of learning above the value of not failing. That, says Stanford University psychologist Carol Dweck, is a hallmark of what she calls “growth mindset.”

 

‹ Prev