Deep in the Heart of Texas - Collider: The Search for the World's Smallest Particles - Paul Halpern

Collider: The Search for the World's Smallest Particles - Paul Halpern (2009)

Chapter 7. Deep in the Heart of Texas

The Rise and Fall of the
Superconducting Super Collider

It was a tragedy, a catastrophe, a scientific Titanic… . The
Superconducting Super Collider is … gone forever.

—HERMAN WOUK, A HOLE IN TEXAS (2004)

In the heart of cowboy country, about thirty miles south of Dallas, the town of Waxahachie bears a brutal scar. Where ninety homes once stood among nearby farmland, 135 acres of land lie arid and abandoned. Pocking the soil are seventeen shafts, hundreds of feet deep, now plugged up and useless. A fourteen-mile tunnel, a masterpiece of engineering that now leads absolutely nowhere, forms an arc-shaped gash underground—believed to be slowly filling with water. Less tangible, but even more indelible, wounds have been the injury to the Texas economy and the damage to the American scientific community’s morale.

No one imagined, when the Superconducting Super Collider (SSC) project was in the planning stages, that its outcome would be so dreadful. Ironically, it came into being through the demise of another ill-fated project, ISABELLE. At the July 1983 meeting that led to the severing of ISABELLE’s lifeline, members of the same federal panel recommended constructing a much higher energy machine. They proposed pursuing a suggestion by Leon Lederman—put forth the previous year at a high-energy workshop in Snowmass, Colorado, for a colossal device with superconducting magnets set in a flat, unpopulated region of the United States. Given CERN’s discovery of the weak boson and Fermilab’s announcement that it had broken accelerator energy records, it was time to think big.

For particle physicists, building bigger and better machines makes perfect sense. What are land and money compared to the benefits of unlocking nature’s deepest secrets? Sure there is always a risk that billions of dollars and years of effort would be wasted, but if we didn’t try we would never be able to complete our understanding of the universe. Yet, as certain critics argue, in many other fields of physics (and science in general), important experiments are conducted at far less cost. Some of these have yielded vital societal benefits—for example, the discovery of transistors revolutionized the electronics industry. Is it then worth it to pursue multibillion-dollar projects? The debate over “big science” versus less expensive, “desk-top” experimentation would play a major role in the fate of the SSC.

Envisioned as a 20 TeV collider—much higher energy than either the Super Proton Synchrotron (SPS) or the Tevatron—the SSC was originally known as the “Desertron.” According to historians Lillian Hoddeson and Adrienne W. Kolb, there were two main reasons for this designation: “It was thought to require a large expanse available only in the desert, and as it was to penetrate the bleak ‘desert’ of physical processes described by the theory of grand unification.”1

The fear was that the newly opened Tevatron would not prove powerful enough to bridge the gap between the final ingredients for the Standard Model and any interesting new physics that lay beyond. No one knew then—and we still don’t know—if the Tevatron has the capacity to find the Higgs, let alone even higher energy particles predicted by various Grand Unified Theories and supersymmetric theories. It would be immensely frustrating if elegant new unification theories, linking the electroweak and strong forces in mathematically satisfying ways, could never be physically tested. The frustration would be akin to discerning life on a distant planet but not being able to reach it. Spanning the broad expanse between imagination and investigation would surely require unprecedented technology.

What if the Europeans developed the next generation of colliders themselves, leaving the home of the brave shaking in its boots? Lederman correctly surmised that CERN’s seventeen-mile Large Electron-Positron Collider tunnel would be a tempting place to insert a ring of superconducting magnets and create a thunderous proton collider. Best get started on a native machine, American physicists realized, before the Old Country helped itself to a buffet of Nobel Prizes.

In December 1983, the Department of Energy formed a planning group, called the Reference Designs Study (RDS), drawing talented individuals, including directors and staff, from the major accelerator labs in the United States. Representatives from Fermilab, Stanford Linear Accelerator Center, Brookhaven, and Cornell ironed out the preliminary details of how one of the most ambitious scientific projects of all time would be implemented. To be fair to all, the meetings rotated from lab to lab. Cornell physicist Maury Tigner, one of Wilson’s former students, headed the group.

By April 1984, the RDS had produced three different options for a proton-proton collider design with 20 TeV of energy per beam and a luminosity of about 1034 (1 followed by 34 zeroes) particles per square inch per second. That’s about 100,000 times more intense than the SPS beams that spawned the W and Z bosons—like packing the entire population of Topeka, Kansas, into a subway turnstile intended to fit a single person. To accomplish this tight squeeze, each option involved a distinct superconducting magnet design—a strong one with a giant iron yoke and a superconducting coil, a relatively weak one with even more iron (called superferric and producing a smoother magnetic field), and a third, medium-strength one that was iron-free (perhaps they already had too many irons in the fire). In a report to the Department of Energy (DOE), the team concluded that all three designs were feasible and recommended further study.

The next phase in planning the SSC took place through a team called the Central Design Group (CDG), situated at Lawrence Berkeley National Laboratory and also led by Tigner. Several different labs, including the Berkeley Lab, Brookhaven, Fermilab, and the Texas Accelerator Center, a research institute established by collider expert Peter McIntyre, were recruited to investigate the various superconducting magnet designs. The majority of the group supported the strong-field magnet that would require a ring about fifty-two miles in circumference, as opposed to the weak-field design that would necessitate a ring up to a hundred miles long. It would be hard enough to find enough land to support the smaller ring.

By late 1986, the SSC design had been submitted to the DOE and the major labs were united in their support. Because of the projected multibillion-dollar cost for the project, to move forward the approval of President Ronald Reagan, the American leader at the time, was required. On the face of it, that would seem to be a hard sell, given that the federal budget was already stretched through various military and scientific programs. Expensive projects at the time included the Strategic Defense Initiative (more commonly known as “Star Wars”) for developing missile-interception systems and a program to establish an international space station. How could the SSC carve out its own piece of a shrinking pie?

Luckily for the prospects of gaining approval, if there was a frontier to be conquered or an enemy to be defeated, Reagan believed in taking major risks. Famously, long before becoming a politician, he played the role of a football player for an underdog team in the film Knute Rockne, All American. Like his character George Gipp, better known as “the Gipper,” optimism and determination steered his judgments. The SSC was a project on the scientific frontier that represented a means of remaining competitive with Europe. That was the angle its proponents emphasized to win over Reagan’s support.

A DOE official asked Lederman if he and his staff could produce a ten-minute video for the president about the questions in high-energy physics the SSC could potentially resolve. To appeal to Reagan’s cowboy spirit, Lederman decided to emphasize the frontier aspects of the research. In the video, an actor playing a curious judge visits a lab and asks physicists questions about their work. At the end he remarks that although he finds such research hard to comprehend, he appreciates the spirit behind it, which “reminds [him] of what it must have been like to explore the West.”2

Apparently the SSC advocates’ arguments were persuasive, because Reagan was deeply impressed. Members of his cabinet, concerned about the fallout from the project’s asteroidlike impact on the budget, ardently tried to intercept it. With all of their strategic defenses, however, they could not shoot it down.

At Reagan’s January 1987 announcement that he would support the SSC, he took out an index card and recited the credo of writer Jack London:

I would rather be ashes than dust,
I would rather my spark should burn out in a brilliant
blaze,
Than it should be stifled in dry rot.
I would rather be a superb meteor,
With every atom of me in magnificent glow,
Than a sleepy and permanent planet.3

After reading the credo, Reagan mentioned that football player Ken Stabler, known for coming from behind for victory, was once asked to explain what it meant. The quarterback distilled its sentiments into the motto, “Throw deep!”4 Surely the erstwhile “Gipper”-turned-president would aim no lower. The SSC would move ahead, if he could help it.

At a White House ceremony the following year, Reagan heralded the value of the SSC, saying, “The Superconducting Super Collider is the doorway to that new world of quantum change, of quantum progress for science and for our economy.”5

Following Reagan’s endorsement came the battle for congressional approval, which would not be so easy. Many members of Congress balked at the idea of the United States carrying the ball alone. Consequently, the SSC was marketed as an international enterprise, involving Japan, Taiwan, South Korea, various European nations, and others. Newsweek described selling the project this way: “The DOE promised from the start that other nations would help bankroll the SSC. That pledge has greased the project’s way through many sticky budgetary hearings.”6

Significant international support was hard to come by, however. The Europeans in particular were naturally more interested in seeing CERN succeed than in supporting an American enterprise. That’s around the time the Large Hadron Collider (LHC) project was first proposed—clearly a higher priority for Europe.

A New York Times editorial on May 20, 1988, argued that any American funding toward the SSC would be better spent contributing toward the LHC instead:

This is a tempting but dangerous initiative because funds to pay for it almost certainly would be stripped from other physics research… . The field is of high intellectual interest, and it would be a sad day if the United States did not remain a major player. But European physicists have shown how an existing collider ring at Geneva could be upgraded to within probable reach of the Higgs boson. Buying into the European ring would be cheaper.7

By mid-to-late 1988, Congress had allocated $200 million toward the SSC. Proceeding with caution, it mandated that the money could be used only for planning and site selection. Because an election was imminent, funds for the actual construction would be left to the consideration of the next administration.

Once Congress offered a kind of yellow light for the project to move forward cautiously, the politics behind the venture became even more intense. Many depressed regions salivated over the potential for jobs. Which state would be lucky enough to acquire what the Times called “a $6 billion plum”?8

To be fair, the competition to find a suitable site was judged by a committee established by the respected and politically neutral National Academy of Sciences and National Academy of Engineering. Of the forty-three proposals submitted from twenty-five states, including seven from the state of Texas alone, the committee narrowed them down to seven. The Lone Star State was clearly the most eager; its government established the Texas National Research Laboratory Commission (TNRLC) well in advance and sweetened the pie by offering $1 billion toward the project from its own budget. What better place to think big than in Texas? Its submission was so hefty—rumored to weigh tons—it was hoisted to the DOE office by truck.9 In November 1988, after considering the relative merits of the finalists, the DOE announced that Ellis County, Texas—a flat, chalky prairie region with little vegetation except for grasses, shrubs, mesquite, and cactus—was the winning location.

By odd coincidence, the site location announcement occurred just two days after a Texan—then Vice President George Herbert Walker Bush—was elected president. The DOE selection group assured the public that politics played no role in the decision. Through a spokesperson, Bush asserted that he had no involvement in the process and that he found out about the choice when everyone else did.10

With the choice of a completely undeveloped site in Texas, many supporters of Fermilab, which had also bid, were left fuming. Fermilab already had much of the infrastructure in place to support an expanded mission; a new location would require starting from scratch with all new staff, buildings, and equipment. In particular, the Tevatron could have been adapted, like the SPS at CERN, to serve as a preaccelerator for the protons entering the collider. “If it was at Fermilab, it would have existed now,”11 said Brookhaven physicist Venetios Polychronakos, who was involved in planning an SSC experiment.

Some feared a brain drain from Fermilab, with top researchers finding new positions in Texas. Lederman, who had a key stake in both institutions as Fermilab’s director and one of the SSC’s original proposers, expressed mixed feelings. He anticipated a “certain loss of prestige” for his own lab.12 However, he expected that it would remain a premier research facility—at least during the years when the SSC was under construction.

Soon after its bid was accepted, the Texas state government, through the TNRLC, assembled a parcel of almost seventeen thousand acres near Waxahachie. For the land of Pecos Bill, that was just a mere pasture. The TNRLC also arranged for environmental studies and administered a research and development program to support the lab. The state’s commitment to the project remained steadfast until the end.

The federal commitment to the project was murkier, as there were many clashing forces involved. Congress, the DOE, and research physicists themselves each had different interests, which were sometimes at odds. Although Tigner’s CDG performed the bulk of the original planning for the project, when it came time to move forward with constructing the collider and setting up the lab, the group was bypassed in favor of the Universities Research Association (URA)—a consortium that managed Fermilab and with which the DOE felt comfortable working. Tigner was seen as a “cowboy in the Wilson tradition” 13 and as potentially having difficulty bending to the demands of Congress and the DOE. Thus, the URA instead chose a relative newcomer, Harvard physicist Roy Schwitters, to become lab director. After serving briefly under Schwitters as deputy director, Tigner stepped down in February 1989 and returned to Cornell. Sadly, with his resignation, the construction of the SSC would need to proceed without his vital technical experience, including the five years he spent helping design the accelerator.

As scientific historian Michael Riordan has pointed out, Schwitters and the URA sharply departed from prior practice by integrating private industrial contractors into the decision-making process. 14 Before the SSC, accelerator laboratories were planned solely by research physicists—who would employ technicians if needed for particular tasks. As we’ve seen, for instance, Wilson designed almost all aspects of Fermilab himself. Schwitters had a different philosophy, involving industrial representatives as well as academics to solve the engineering challenges associated with building what would have been the world’s most formidable scientific apparatus. Among these were corporations heavily involved in the defense and aerospace industries, and for whom it was their first exposure to high-energy physics. Many of the industrial workers switched jobs because of military cutbacks due to the end of the cold war. Because they were used to a certain mind-set, the lab assumed some of the secretive aspects of a defense institute. Moreover, some of the established physicists felt that they couldn’t handle the scientific demands. These factors, as Riordan noted, created a clash of cultures that alienated many of the experienced researchers and made it hard to recruit new ones.

Despite these internal tensions, during the first Bush administration, the SSC benefited from strong federal support. Familiar with the Texas landscape due to his years as a businessman and politician in that state, President Bush considered the SSC a national scientific priority and pressed Congress each year to fund it. The DOE distributed particular tasks such as building and testing detectors throughout almost every state, offering politicians around the county further incentive to favor it.

The program even survived a doubling of its estimated completion cost to a whopping $8 billion, announced in early 1990. This came about when engineering studies forced a major redesign of the project due to issues with the magnets and other concerns.

Superconducting magnets are in general extremely delicate instruments. The stronger a magnet’s field, the greater its internal forces and the higher the chance that its coil and other parts will subtly oscillate. Vibrations cause heating, which can ruin the superconducting state and weaken or destroy the magnets. At 6.5 Tesla (the metric unit of magnetism)—more than 50 percent stronger than the Tevatron’s field—the SSC’s magnets were very much at risk. To minimize tiny movements, the magnets included carefully placed steel clamps. Getting them to work effectively was a matter of trial and error. In an important preliminary test of twelve magnets, only three passed muster. Designers struggled to improve the performance.

Another question concerned the size of the magnets’ openings. Smaller apertures were cheaper but entailed a greater risk to the streams of protons passing through. Any misalignment could reduce the rate of collisions and sabotage the experiment. Ultimately, after considerable discussion, the SSC administration decided to enlarge the magnet openings to allow more room for error.

Other design changes made at that time included increasing the ring circumference to fifty-four miles and doubling the proton injection energy (the energy protons are accelerated to before they enter the main ring). All of these modifications forced the bill sky high. Although some members of Congress became livid when learning about the huge increase in the cost, the general reaction at that time was to increase oversight rather than close down the project. Construction funds started to flow, and the lab began to take shape starting in the fall of 1990.

As planned, the SSC was to have a succession of accelerators boosting protons to higher and higher energies before they would enter the enormous collider ring. These consisted of a linear accelerator and three synchrotrons of increasing size: the Low Energy Booster, the Medium Energy Booster, and the High Energy Booster. Tunnels for the linear accelerator and the smallest synchrotron were the first to be excavated.

Boxy structures, to support future operations underground, sprang up like prairie grass along the flat terrain, including the Magnet Development Lab, the Magnet Test Lab, and the Accelerator Systems String Test Center. These were facilities for designing, building, and testing the various types of superconducting magnets needed for the project. Two large companies, General Dynamics and Westinghouse, took on the task of building the thousands of dipole magnets that would steer the protons, like roped-in cattle, around what would have been the largest underground rodeo in the world.

Meanwhile, drawn by the prestige of contributing to a kind of Manhattan Project for particle physics, or simply finding a good-paying position, more than two thousand workers relocated to Texas. The lure of potentially finding the Higgs boson or supersymmetric companion particles enticed many an adventurous physicist to venture south of glittery Dallas and try his or her luck with collider roulette. For researchers who already had thriving careers, it was a significant gamble. Some took leave from their full-time positions; others gave up their old jobs completely with hopes of starting anew.

In the tradition of colliders supporting a pair of major detectors, with collaborations lined up behind these, two groups’ proposals were approved for the SSC. The first, called the SDC (Solenoidal Detector Collaboration), involved almost a thousand researchers from more than a hundred institutions worldwide and was headed by George Trilling of the Berkeley Lab. Its general-purpose detector was designed to stand eight stories high, weigh twenty-six thousand tons, and cost $500 million. The target date for it to start collecting data was fall 1999—offering hope of finding the Higgs before the toll of the millennial bells.

The second group, called GEM (Gammas, Electrons, and Muons), was led by Barry Barish of Caltech, an accomplished experimentalist with a stately beard and shoulder-length silver hair, along with liquid argon calorimetry coinventor William Willis of Columbia University, and a humongous cadre of researchers. Their project involved a detector specially fashioned for pinpoint measurements of electrons, photons, and muons. GEM was supposed to be located around the ring from the SDC at a different intersection point, collecting data independently, like competing newspapers housed in separate city offices.

Unfortunately, neither of these detectors ever had a chance to taste flavorful particles. As the SSC project rolled through the early 1990s, it accumulated more and more opposition—not just from politicians dismayed that it would break the budget but also from fellow physicists in fields other than high energy. Most branches of experimental physics don’t require $8 billion budgets, yet can still yield groundbreaking results.

Take for example high-temperature superconductivity. In the 1980s, Swiss physicist Karl Müller and German physicist Johannes Bednorz, working with reasonably priced materials in the modestly sized (compared to CERN or Fermilab at least) complex of IBM’s Zurich Research Laboratory, revolutionized physics with their discovery of a ceramic compound that could conduct electricity perfectly at temperatures higher than previously known superconductors. Other experiments in various labs, including work by Paul Chu at the University of Houston, turned up substances with even higher transition temperatures. Although these ceramic superconductors still need to be quite cold, some maintain their properties above the temperature of liquid nitrogen. Immersing a material in liquid nitrogen is far cheaper than the drastic methods used to create the near-absolute-zero superconductors once believed to be the only types. Therefore not only did Müller and Bednorz’s finding come with a much cheaper price tag than, say, the top quark, it also led to cost saving for future research and the potential for more widespread applications of superconductivity.

Because discoveries related to material properties bear more directly on people’s lives than does high-energy physics, many researchers in these fields, such as Cornell physicist Neil Ashcroft, have argued that they deserve at least as much support. “Things are out of whack,” he said. “Condensed-matter physics is at the heart of modern technology, of computer chips, of all the electronic gadgets behind the new industrial order. Yet relative to the big projects, it’s neglected.”15

Another leading critic of “big science,” who was skeptical about channeling so much funding into the Super Collider, was Arno Penzias, codiscoverer of the cosmic microwave background. Penzias said, “One of the big arguments for the S.S.C. is that it will inspire public interest in science and attract young people to the field. But if we can’t educate them properly because we’ve spent our money on big machines instead of universities, where’s the point? As a nation we must take a new look at our scientific priorities and ask ourselves what we really want.”16

On the other hand, who could anticipate what would have been the long-term spin-offs of the SSC? In the past, some discoveries that seemed very theoretical at the time, such as nuclear magnetic resonance, have ended up saving countless lives through enhanced techniques for medical imaging and treatment. But since nobody had a crystal ball for the SSC and its potential applications, its critics painted it as just big and expensive.

The rising crescendo of arguments against “big science” and in support of smaller, less expensive projects jived well with growing congressional sentiments that the SSC was getting out of hand. Given that Congress was promised that substantial foreign contributions would fill out the SSC’s budget, when these failed to materialize, many members were understandably miffed. Some didn’t think that Schwitters and the DOE under Secretary of Energy James Watkins were managing the project effectively. Still, it came as a surprise when in June 1992 the House of Representatives voted 232-181 in favor of a budget amendment that would end the project.17 Only the Senate’s support for the SSC temporarily kept it alive.

In the spirit of former senator William Proxmire’s “Golden Fleece Awards” for alleged government boondoggles, many of those who favored terminating the SSC painted it as a wasteful endeavor that would benefit only a small group of eggheads. In times of tight budgets, they wondered, why channel billions of dollars into crashing particles together to validate theories rather than, say, blasting away at the all-too-real federal deficit behemoth?

“Voting against the SSC became at some point a symbol of fiscal responsibility,” said its then associate director Raphael Kasper, who is currently vice president of research at Columbia. “Here was an expensive project that you could vote against.”18

In January 1993, Bill Clinton succeeded Bush as president. Without the Texas connection, a key strand of the SSC’s support dissolved. Although Clinton indicated that he backed the project, particularly in a June letter to the House Appropriations Committee, he advocated extending the time line for three extra years to reduce the annual impact on the federal budget. Postponing the SSC’s targeted opening date (to 2003) made it seem an even riskier venture, however, because it could well have been obsolete by the time it went on line. What if the Tevatron had found the Higgs boson by then?

Once the collider lab’s anticipated costs rose to approximately $10 billion, largely because of the pushing back of its schedule, it was only a matter of time before an increasingly frugal Congress signed a do-not-resuscitate order. A House of Representatives vote on October 19, 1993, denied by a two-to-one margin a funding bill that would have supported further construction. Instead, the SSC’s annual appropriation was directed to moth-ball the part of the facility that had already been built. By then $2 billion had already been spent and more than one-quarter of the project was complete—all for naught. The tangible result of a decade of planning and hard work would just be boarded up and shrouded in dirt. Requiescat in pace.

The cancellation of the SSC did, in the short term, save federal money. Along with many other cost-cutting measures, the federal budget would be balanced by the end of the decade. (Ironically, in the 2000s, the deficit would skyrocket again, making all of the cost cutting moot!) Yet, what is the long-term price of a national decline in scientific prestige? Skipping the moon landings, eschewing the robotic exploration of Mars, and abstaining from telescopic glimpses at the swirling mists of ancient galaxies would have each cut government expenses, too—while extinguishing the flames of our collective imagination. If it is a choice between science and sustenance, that’s one thing, but surely our society is rich enough to support both. It remains to be seen whether the United States will ever resume its pioneering mantle in high-energy physics. Thus in retrospect, many see the abandonment of what would have been the premier collider in the world as a grave error.

According to Fermilab physicist William John Womersley, “The SSC has cast a very long shadow over high-energy physics and big science in general. We’re still dealing with the legacy.”19

In the aftermath of the closure, those who took the career risk and moved down to Texas for the SSC met with varying degrees of disappointment. Some regrouped, sent out their résumés (or were recruited), and managed to find new positions in other labs or universities. For the experienced physicists, finding an academic position was hard, because not many universities wished to hire at the senior level, and the closure of the SSC reduced the need for professors in the high-energy field. A survey taken one year after the closure found that while 72 percent of those in the SSC’s Physics Research Division had found employment, only 55 percent of those positions were in high-energy physics.20

Other workers, who had laid down deep roots in Texas and didn’t want to leave, either found other types of jobs or simply retired early. A few stayed to help sell off the equipment and assist in attempts to convert the site to alternative uses.

Given all of the time and energy that went into assembling the land, digging the tunnels, and constructing the buildings, it is remarkable that the site has yet to be put to good use. The federal government transferred the property to the state of Texas, which in turn deeded it to Ellis County. For more than fifteen years, the county has tried in vain to market the structures, particularly the former Magnetic Development Laboratory. Like Dickens’s forlorn spinster, Miss Havisham, the relic building is a jilted bride frozen in time with no interested suitors. An agreement to convert it into a distribution center for pharmaceutical products fell through, and informal plans to house an antiterrorism training base never materialized. In 2006, trucking magnate J. B. Hunt’s plans to use it as a data center were abandoned upon his death.21 It did, however, play a background role in a straight-to-video action flick, Universal Soldier II. 22 To mention another has-been, the Norma Desmond of labs finally had its close-up.

Though it’s instructive to ponder what could have been in hypothetical scenarios about alternative choices, in truth physicists can’t afford to wallow in disappointment. An energetic frontier is ripe for exploration and there’s no time for looking back. Leaving the plains and pains of Texas behind, in the late 1990s the American particle physics community regrouped and headed either north to Illinois, for renewed efforts at the Tevatron, or across the ocean to the cantoned land where cubed meat and melted cheese deliciously collide. After all, Geneva, Switzerland, has distinct charms, some of which Lederman described well. Comparing it to Waxahachie, he wrote, “Geneva … has fewer good rib restaurants but more fondue and is easier to spell and pronounce.”23

Humanity’s best chance of finding the Higgs boson and possibly identifying some of the lightest supersymmetric companion particles now rests with the Large Hadron Collider. Though it will crash particles together at lower energies than the SSC was supposed to—14 TeV in total instead of 20 TeV—most theoretical estimates indicate that if the Higgs is out there the LHC will find it. If all goes well, modern physics will soon have cause for celebration.