The Boy Who Played with Fusion

Home > Other > The Boy Who Played with Fusion > Page 23
The Boy Who Played with Fusion Page 23

by Tom Clynes


  * * *

  Heavy Metal Apron

  FOR TAYLOR, SUDDENLY, things were happening. In contrast to Winterberg’s inflexible, theory-based approach, the people Taylor was meeting now actually did things and built things. Within a few days, Phaneuf had cleared a generous work area for Taylor at one end of his own high-tech physics laboratory. Brinsmead brought Taylor to the physics machine shop and introduced him to the department’s engineers, machinists, and technicians, including Wade Cline, a metallurgy and welding whiz who would become a good friend.

  Most afternoons after school, Taylor would walk over to the Leifson Physics Building with his chaperone from the university and descend to the subbasement, where he’d spend the next two hours. On Fridays, when he had just one class at Davidson, he’d spend most of the day at the lab. Phaneuf and Brinsmead checked that Taylor was badged with a dosimeter the whole time, and they outfitted him with a big lead apron, like someone would wear for a dental x-ray, that came down almost to his shoes.

  A physics lab can be a lonely place, and Phaneuf appreciated Taylor’s energetic company—in fact, Phaneuf says, the late afternoons with Taylor quickly became his favorite part of the day. “He’s a happy guy to start with, and why wouldn’t he be? He was doing what he wanted, and a lot of it. He’s a doer, a creative person, and just filled with ideas. It’s a joy to watch him learn, and hear him think and create.”

  Taylor is attracted to anything energetic, and there’s nothing more energetic than a hot-plasma environment. Phaneuf, one of the world’s leading plasma researchers, helped Taylor deepen his understanding of the enigmatic workings of this fourth state of matter while he conducted his own experiments on the snaking, highly complex ionization machine that takes up most of the space in the lab. It’s one of only two in the world, both of which Phaneuf built from scratch.

  Phaneuf used his machine to conduct experiments on how various elements behave when ionized. His research has informed several disciplines and has helped other researchers understand deep-space nuclear reactions by simulating astrophysical hot-plasma environments.

  Manmade low-temperature plasmas light up fluorescent bulbs and neon signs. When electricity flows through tubes containing certain gases, it ionizes and excites the atoms and creates glowing plasma inside the tube. But high-temperature plasmas are rare on and near Earth. Auroras (the northern and southern lights) occur when charged particles from the sun get caught in the Earth’s magnetic fields. Spiraling toward the poles, some of these particles reach the upper atmosphere, where they ionize nitrogen and other elements, forming plasma. As these ions recombine with electrons light is emitted, giving us spectacular light shows.

  While natural plasmas are scarce on and near our planet, about 99 percent of the known matter in the universe exists in the plasma state. For instance, stars are big balls of plasma that host nuclear fusion reactions. Taylor knew that he’d need to deepen his understanding of high-temperature plasma to achieve fusion; under Phaneuf’s tutelage he soon became obsessed with this temperamental state of matter.

  The Farnsworth-Hirsch fusor is an ingenious device. “Think of it as a Polish firing squad,” Phaneuf says with a laugh. “You’ve got ions in a circle shooting inward. If all goes well some will meet head-on in the middle, violently enough for the nuclei to fuse together, releasing high-energy neutrons.”

  Near the middle of the spherical reaction chamber is a negatively charged spherical grid. When the fusor is fueled with deuterium gas and electrified, the high-voltage current strips electrons off the deuterium atoms, converting them to positively charged atomic nuclei (ions) that fly toward the negatively charged inner cage. When the pressure, voltage, and fuel mix are just right, a glowing core of superheated plasma forms inside the cage. Mimicking the core of the sun, its ultrahigh density serves as an attractive target for the converging ions.

  If the converging ions miss the cage and one another, they fly out the other side. But as they move outward, they come under the influence of the inward electrostatic force, which directs them back toward the center, where they have the chance to meet again. The majority of impacts don’t result in fusion, but some are energetic enough to overcome the electromagnetic forces that push them apart. Then the strong nuclear force takes over and the nuclei fuse, releasing their energy in the form of neutrons.

  In a fusor, the fusion reactions are driven by particle velocity more than heat, so little energy is lost through light and other radiation. The lack of magnetic fields also lowers energy losses. But a fusor’s major limiting factor is the grid itself, which robs charged particles of their energy when they strike it. This hobbles the fusor’s capacity to produce power and wastes the energy invested in ionizing and accelerating the particles. The strikes also cool the plasma down and can damage the fine-mesh grid itself.

  A Farnsworth-Hirsch fusor is far easier to describe than it is to build. To get it right, Taylor needed to design and construct a system with such precision that it could maintain vacuum pressure several orders of magnitude lower than the vacuum of outer space. He needed to precisely concentrate up to 100,000 volts of electricity and coax his plasma core up to about 580 million degrees within an environment robust enough to sustain and survive those torturous conditions. And to verify his results, he needed to develop a detection system that could accurately differentiate neutrons from other types of radiation and document their presence convincingly enough to overcome the skepticism of the Neutron Club gatekeepers at the Fusor.net forum. Thinking ahead to the process of data acquisition, Phaneuf helped Taylor set up a LabVIEW software system and got him started on coding the system’s data-collection functions.

  Phaneuf had helped to get Taylor’s nuclear fusion project moving by providing the space, critiquing Taylor’s designs, and cutting through bureaucracy. As work progressed, Brinsmead grew more involved, to the point that he was there with Taylor most afternoons. Their friendship grew as they brainstormed design ideas, scrounged parts, and “conned the machinists to help Taylor do some work on it,” as Brinsmead put.

  “Bill was pretty much perfect for Taylor,” Phaneuf says. “He had the right personality for mentoring. He made it fun, but he could also say, ‘Hey, Taylor, wait a minute; before you do that, you’d better check this.’ That’s especially important in this field, because if you get excited or tired and make a mistake, it can be lethal. Taylor is excitable and somewhat impulsive by nature, so we really worked with him to get him to stop and assess a situation before taking action.”

  Mentorship can be the key to keeping very bright kids headed in the right direction. The vast majority of those who go on to illustrious careers or discoveries say that having a mentor was an important part of their development. And yet, meaningful mentorship is largely absent in the education of not only the gifted and talented, but all children. That’s partly because the process of setting up effective one-on-one relationships is hard to institutionalize. Finding the right person for a student, someone who has expertise and time and the willingness to take a personal interest, can take a great deal of energy and effort. As resources for education have been cut and workloads have increased, it has become harder for educators and institutions to be good mentorship partners.

  That shouldn’t let schools off the hook, advocates for gifted and talented children insist. “All the AP and summer programs in the world are not enough,” says researcher David Lubinski. “Kids respond much better to a mentor; those one-on-one relationships are extremely important and extremely effective if you get the match right.”

  Phaneuf’s and Brinsmead’s hands-on brand of science was a good match for Taylor. “You can’t buy the stuff in my lab, just like you can’t buy a fusor,” Phaneuf says. “We have to roll up our sleeves and build practically everything from scratch. So what Taylor was trying to do fit right into the style of what we were doing.”

  “Taylor was really lucky to find Ron and Bill,” says Elizabeth Walenta, the science teacher. “They were able to see this
tiny fourteen-year-old for who he was and what he could do, and they could help him do it safely.”

  But it was Taylor’s persistence and initiative, not luck, that got him the guidance he needed after he was shot down by Winterberg. “The right mentoring is essential,” says Simonton, “but it’s not always easy to obtain. A significant part of talent is the willingness to seek out the best mentors, even when rejection is a possibility.”

  “I wish UNR would make it easier and get more involved in mentorship with students,” says Walenta. “A lot of kids want to do that, and it can really work for a student like Taylor, who is motivated and gregarious and affable. But others are not as persistent. I send some kids over to the university, and they get lost and hit a wall and give up.”

  “Taylor, you’re definitely a hands-on guy,” Brinsmead told Taylor one day as they inventoried the hundred and fifty or so parts Taylor had brought with him from Arkansas. “Not too many kids are like that anymore; they don’t want to get their hands dirty and try things like you do.

  “But the problem is,” Brinsmead continued, “the stuff you want to build is so incredibly expensive.”

  According to amateur fusion pioneer Tom Ligon, acquiring the parts to build a working fusor is the second-biggest stumbling block, after acquiring the expertise. “Unfortunately,” says Ligon, “most of the people who start on a project like this are too poor to leap in and build something quickly. The people who stick with it are typically above normal in terms of their motivation and tenacity.”

  As it turned out, very few of the parts Taylor had brought from Arkansas would prove up to the task. But when it came to scrounging, Taylor had one of the best mentors imaginable. Brinsmead downplays his influence. “I may have taught him a few tricks, but let’s face it: there isn’t a lot that Taylor can’t talk someone into—or out of.”

  Taylor made the rounds of his growing group of friends and supporters in the physics department and beyond, scaring up more ConFlat fittings, finding a source of finer-gauge tungsten for the fusor’s grid, taking an electrical control panel off the hands of a professor who no longer needed it. He and Kenneth and Brinsmead drove out to a closed bank to confirm rumors of a video camera left behind, and Kenneth used his Rotary Club connections to track down someone who could provide permission for them to take the camera. They visited one of Brinsmead’s old surplus haunts and found an IBM disk-drive frame from the late 1970s that could support the machine.

  Taylor had a head start with his stainless-steel chamber, but locating other critical parts at prices he could afford was a major challenge. It would be crucial, as Taylor had already learned, to create and sustain a high vacuum. He had started with refrigerator compressors wired to run backward, then advanced to single-stage vacuum pumps, then to the oil-diffusion pump from the astronaut’s mass spectrometer. But both Brinsmead and Phaneuf doubted that the oil-diffusion pump was up to the job. Taylor, making the rounds of the physics department, heard that a professor had purchased a two-stage turbo-molecular pump for an experiment but a grad student had let it roll around in the back of a truck and some of the parts had fallen off, and now it was inoperable.

  “That was a twelve-thousand-dollar bearing-stabilized turbo pump,” Phaneuf marvels, “and somehow, Taylor ended up with it because someone didn’t want to bother to fix it.” Taylor was able to buy or build all the missing parts for the turbo pump, which could spin at jet-engine speeds. Once he got it working, he teamed it with the diffusion pump to create a two-stage system strong enough to reduce the pressure to a workable level.

  Even though Taylor eventually found his way around Winterberg, the encounter had been influential, forming the roots of Taylor’s growing distaste for theoretical physics. It would cement his devotion to hands-on invention as he became less interested in the paths of the Einsteins and Heisenbergs and more focused on putting their discoveries to work in the manner of Tesla and Farnsworth. “I want to apply the physics theory that’s been discovered,” he’d say, “and use it to create something new to solve real-world problems.”

  As a result of the clash with Winterberg—who had told Taylor to master calculus and then work his way up through the theoretical physics courses before attempting anything hands-on—Taylor “dragged in a fair bit of defiance” to his math classes, according to teacher Darren Ripley. Taylor doesn’t disagree. “I had a chip on my shoulder,” he says, “and some of it is still there.” In New Mexico, Taylor told Willis that he didn’t believe calculus was “all that useful.” When Willis raised a skeptical eyebrow, Taylor responded unyieldingly: “Carl, I don’t want to be a theoretician!”

  And yet, without a thorough understanding of physics theory, it would be impossible for Taylor to build a fusion reactor that could mimic the power-generation process of the sun, impossible to understand the behavior of the subatomic particles that would, he hoped, collide and fuse in his reactor, releasing their neutrons. Winterberg was right: to be a physicist, whether theoretical or applied, Taylor would need to stand on a solid scaffolding of high-level mathematics.

  Ripley’s job—to help him build that scaffolding—was made tougher by the fact that Taylor, like most of his incoming students, had been “doing math in a way in which they don’t know where the concepts come from.” It’s a criticism I would hear from a lot of experts in STEM education: Mathematics continues to be taught as if it has no history and very little future application to what kids see as relevant to their own interests and career plans. “It’s usually just drill and kill,” says Ripley, “and that’s what kills interest in math.”

  But Ripley is a talented teacher, and his small class sizes allowed him to know his students well enough to make the material applicable to each student’s particular passion. Ripley also understood Taylor’s dominant learning style—he makes sense of things by talking about them—and so he would often turn the discussion over to Taylor to connect a mathematical concept to some phenomenon of radiation.

  “Tay, why don’t you tell us why this is useful?” Ripley would say. “Then,” Ripley tells me, “once he got going, it could get pretty entertaining, and you could see the light go on. Suddenly, it’s relevant to him. You could see new pathways forming, wheels turning, deeper thought processes.”

  Taylor and Phaneuf both agree that Taylor got a lot out of Phaneuf’s upper-division nuclear physics class—“although,” says Phaneuf, “some of the math was beyond him.” Phaneuf suggested that Taylor take UNR’s nuclear chemistry class with Dennis Moltz, a guest professor from Berkeley.

  According to Moltz, whenever he asked the class a question, Taylor would blurt out the answer. At Davidson—where Taylor’s peers shared his intellectual intensity—that wasn’t a big deal. But in the university environment, it didn’t go over so well. Moltz had to take Taylor aside and ask him to restrain himself.

  Taylor was exhibiting another common trait of highly gifted children: intellectual overexcitability, or intellectual OE. According to psychologists, those high in intellectual OE are intensely curious and usually able to concentrate, engage in prolonged intellectual effort, and tenaciously problem-solve when they choose. But their extreme eagerness can cause social difficulties when they become so excited that they interrupt at inappropriate times or grow impatient with others who can’t keep up with their intellectual pace. Polish psychologist and physician Kazimierz Dąbrowski, whose studies of gifted schoolchildren led to the theory of intellectual OE, says that people with high intellectual OE tend to have powerfully creative perceptions of the world.

  But not all creativity is created equal. Mihály Csíkszentmihályi and other psychologists and creativity researchers distinguish between what they call big-C and little-c creativity. Little-c creativity includes garden-variety problem solving and the ability to adapt to change. Big-C creativity, which is far more rare, occurs when a person solves a problem or creates something original that has a major impact on other people.

  “At the little-c level, creativity implies bas
ic functionality,” says Simonton. “And at the big-C level, it’s something that we give Pulitzer and Nobel Prizes for.”

  A former president of Caltech contended that one truly excellent scientist is more valuable than a thousand very good scientists. “Indeed, it seems unlikely that 1,000 average scientists could have produced the General Theory of Relativity,” wrote aerospace pioneer Norman Augustine, “no matter how much time they were allocated. Nor could 1,000 average writers have produced Shakespeare’s works. Nor could 1,000 composers have created Beethoven’s music.”

  One challenge for gifted kids, says Ellen Winner, is making the leap from being a gifted child to being an influential creator. “The skill of being a child prodigy is not the same skill as being a real big-C creator. A child may be extremely good at mastering something, but most kids doing calculus at eight aren’t necessarily creators. That’s a very different kind of mind. Most prodigies don’t become creative geniuses, they become experts. Only some manage to move beyond that to a more reflective sense of direction.”

  23

  * * *

  Birth of a Star

  SEVERAL MONTHS INTO his fusion project, Taylor was still missing a few parts, but he had, thus far, spent less than two thousand dollars on his fusor, even though it contained tens of thousands of dollars’ worth of precision parts. Some had been custom made in the physics department’s state-of-the-art machine shop, others scrounged from equipment rooms, surplus suppliers, UNR professors, and friends and colleagues around the country.

  But some of the more specialized (and critical) parts were proving harder to find. The electrical insulator Taylor had finagled didn’t have enough insulation to keep it from ionizing the air around it (much like high-voltage power lines do, a phenomenon that’s often noticeable in wet weather). Whenever he cranked the voltage up, the power supply would start arcing and sputtering.

 

‹ Prev