Bad Science: The Short Life and Weird Times of Cold Fusion [1 ed.] 0394584562, 9780394584560

A science journalist brings to life one of the greatest scientific frauds of our times with the story of the two obscure

191 39 33MB

English Pages 503 [534] Year 1993

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Bad Science: The Short Life and Weird Times of Cold Fusion [1 ed.]
 0394584562, 9780394584560

Citation preview

a ~.*

ALSo

BY

GARY

TAUBES

Nobel Dreams

BAD SCIENCE

Digitized by the Internet Archive in 2019 with funding from Kahle/Austin Foundation

https://archive.org/details/badscienceshortlO000taub

BAD SCIENCE The Short Life and Weird Times of Cold Fusion

GARY

TAUBES

Thomas J. Bata Libror,

TRENT UNIVERSi: y

PETERBOROUGH, ONTARIC.

‘a RANDOM HOUSE NEW YORK

NED

Copyright © 1993 by Gary Taubes All rights reserved under International and Pan-American Copyright Convention. Published in the United States by Random House, Inc., New York.

Library of Congress Cataloging-in-Publication Data Taubes, Gary Bad science: the short life and weird times of cold fusion / Gary Taubes.

py |cms ISBN 0-394-58456-2 1. Cold fusion. I. Title.

QC791.775.C64T38 1993 539.7'64—dc20 91-52693 Manufactured in the United States of America 24689753 First Edition

For my parents,

Ernest Paul Taubes and

Zelda Taubes (April 20, 1920, to May 30, 1991)

x

9 Fie ay

a

The rare genius with a flair for research will not benefit from instruction in the methods of research, but most would-be

research workers are not geniuses, and some guidance as to how to go about research should help them to become productive earlier than they would if left to find these things out for themselves by the wasteful method of personal experience. WILLIAM BEVERIDGE, The Art of Scientific Investigation The first principle is that you must not fool yourself and you’re the easiest person to fool. RICHARD FEYNMAN, “Surely You’re Joking, Mr. Feynman!’’

~

.

ont tenes) Seve He semen) +o sich @ fire coarag emeser gieb! oer beeen deal iaxedtcn by olendieee oct of eomieaeranint 33 an oon mhnky Gites Bap biota

4 ee

Agi

Ro ieseet

wang wine 144 os sed? qled binods Snesees Ipods a_i and Sisco sqrt aca e bark ot teaktt hace qd? wsats wites sviponh regs tantileg hs foiteth Masteew ob yd seviewnedt act iatlanaed

3 IN ESN Pl LAO PEAR

movies Agate

yw

bets Miscutoy eps tes Guerre

_ iawones asennth

lad ab eghaing ter oT

somt espesenssies

sywor 7 ‘Seamanysd hb gabdsy

nih

AUTHOR’S

NOTE

The cold-fusion episode teaches two lessons that can be applied as meaningfully to journalism as to science: 1. Do your research, because nothing is as simple as it seems. 2. Make sure you’ve got the story right before you publish. With these points in mind, over 260 persons were interviewed for this book between March

27, 1989, and November

1992. Some of the

interviews were as brief as a five-minute phone conversation; most took hours or even days. Quite a few sources kindly consented to repeated telephone calls and time-consuming interviews for the duration of the three-year period. A list of those interviewed can be found at the back of the book. Conspicuously absent from this list are Stan Pons and Martin Fleischmann, who refused numerous requests for interviews. Martin Fleischmann, however, was kind enough on several occasions to answer

a few brief questions before having second thoughts. The book was read in draft form and corrections suggested by Allen Bard, Tim Fitzpatrick, Richard Garwin, William Happer, John Huixi

xii

BHAUTHOR'S

NOTE

zenga, Steven Koonin,

Del Lawson,

Nathan

Lewis, Charles Martin,

Hugo Rossi, and Michael Salamon. Any errors remaining in either fact or form, however, are mine alone.

A regrettable side effect of interviewing so many individuals on such a small, however perverse, subject is that one invariably’ ends up with considerably more information than can be accommodated in a book of reasonable size. My editors, justifiably, refused to publish 700-plus pages on cold fusion. As a compromise, they allowed me to transfer much of the less pertinent information to the end notes. This section of the book provides a second or even third level of information, speculation, perspective, and humor into which readers can delve if they choose.

CONTENTS

Author’s Note Prologue: The Press Conference BOOK I: Delusion Is the Better Part of Grandeur Chapter 1: The Meltdown Chapter 2: The Competition Chapter 3: Autumn 1988 Chapter 4: January and February 1989 Chapter 5: February 26 to March 15, 1989 Chapter 6: March 16 to March 23, 1989: The Wager Chapter 7: March 23 and 24, 1989: Afterthought BOOK 11: A Collective Derangement of Minds BOOK 11: The Tail of the Distribution Epilogue Acknowledgments Notes Interviews

Index xiii

GPO,

;

i

oD

i

;

tS

7

e-TRU oC 4

\



r

i.

7

t

.

7 }

7

:

rt

Lay

j

+,

Fo

Wi; 7

~?

a

a,

| °

¥.

ry

PROLOGUE: THE

PRESS

CONFERENCE

GJ March 16, 1989, University of Utah president Chase Peterson

agreed with his scientists and lawyers that it was no longer possible to keep secret the discovery of cold nuclear fusion. As he later explained, the determining factor in his decision was not that his scientists had unambiguous evidence of this remarkable breakthrough, only that such a nuclear reaction might be possible and that they might have created one in their basement laboratory. The research had been the brainchild of B. Stanley Pons, chairman of the Utah chemistry department, and Martin Fleischmann, Pons’s British

collaborator and mentor. They had told Peterson that they had been working together on cold nuclear fusion for five years and that they needed eighteen more months to complete their experimental protocol. But by March 1989, events beyond their lab were dictating the schedule of their research. Time had become a luxury they could not afford. At nearby Brigham Young University, a physicist named Steven Jones had come upon the same remarkable phenomenon. Pons and Fleischmann XV

xvi B PROLOGUE

believed that Jones had not discovered cold fusion independently but had “pirated” the key ideas from them. Six months earlier, the two chemists had submitted a proposal to the

Department of Energy requesting funding for their cold fusion experiments. The DOE had sent the proposal on to Jones for review. Pons and Fleischmann believed that Jones had read their proposal and then run a quick and dirty version of their experiment. Now Jones not only claimed that he had equal right to pursue the research but within a month would disclose that he had discovered cold fusion. On March 16 Pons and Fleischmann had virtually no data to support their experimental conclusion—which is to say, their discovery of cold nuclear fusion—but because of this untimely challenge they had no choice but to make a public announcement and to do it quickly. If not, they would lose credit for what appeared to be one of the great scientific discoveries of all time. The chemists and the university administration agreed to schedule a press conference for March 23, 1989. Until then, secrecy would be paramount. Stan Pons informed his own

researchers of the press conference only two days in advance and swore them to silence on the few details he was willing to divulge. By then they knew that something big was in the works because Pons had seemed even more frenetic than usual. His working hours, which had always been long, had lately verged on twenty-four hours a day. And what the graduate students called “‘administration types” had begun traipsing mysteriously into the back room where the fusion cells were hidden. Pam Fogle, news director at the University of Utah, only began alerting her contacts in the media on March 22. Fogle had been warned by her boss, a fellow with twenty-five years’ experience in public relations, that the discovery was so astonishing that reporters were bound to break any embargo if given half a chance. Following this strategy, Fogle did not send out press releases in advance, fearing they might “fall into the wrong hands.”’ To some reporters—Philip Hilts of The Washin gton Post, for instance—Fogle refused even to divulge what the announ cement would be, only that it was tremendously important and that it would be made at 1:00 p.m. the next day. This sort of secretive argume nt didn’t sway Hilts, so he didn’t make the trip to Salt Lake City. Another of Fogle’s first calls went to Jerry Bishop, the vetera n science reporter of The Wall Street Journal. Fogle needed the Joumal’s coverage, so with Bishop she risked revealing her news—that the Unive rsity of Utah had achieved a sustained nuclear fusion reaction at room temperature for one hundred days. ““That’s ridiculous,” he said, ‘“‘but thanks for letting me know.” Bishop

PROLOGUE

@ xvii

then went off to lunch with a science writer, who suggested that maybe it wasn’t so ridiculous, and he set out to track down the details.

Despite Fogle’s precautions, the CBS News affiliate in Salt Lake City broke the news of cold nuclear fusion that night, and The Wall Street Joumal and the Financial Times of London weighed in the next morning. Rather than diminish interest in the press conference, however, these

leaks further legitimized the university’s announcement. With the news spreading fast, campus police appeared in force at Pons’s basement laboratory on the morning of the twenty-third to guard the discovery of the century. A half-dozen uniformed officers and at least as many plainclothes detectives stationed themselves at strategic points in and around the lab. Security was so tight that when the detectives realized one of the laboratory doors had a window, they ordered the students to cover it with black paper. The graduate students found these actions ludicrous. Inevitably, rumors began spreading throughout the lab and the university at large. One had it that President George Bush would attend the announcement; then it was Margaret Thatcher who was coming. Finally, it was whispered that Vice President Dan Quayle would be there, although he didn’t show either. Considering the magnitude of the discovery, no rumor seemed too outrageous to be rejected out of hand. The second day of spring in Salt Lake City was a balmy one, with a touch of winter lingering in the air. The mountains behind the city and the campus were capped with snow. Many of the faculty and students of the University of Utah would spend the weekend skiing. By 1:00 p.m., the lobby of the Henry B. Eyring Chemistry Building was overflowing. Visitors were packed half a dozen deep by the rear doors. Seven television cameras stood on tripods along the back wall; klieg lights offered a hint of the surreal. The cherubic Jim Brophy, with his white hair and rosy cheeks, vice president for research at the university, began the press conference and made the appropriate introductions. He started with Robert Nesbitt, a dean in the Faculty of Science at the University of Southampton, England. Nesbitt had flown over the previous night to represent, in Peterson’s words, “the international cooperation in scientific exploration” that had made the discovery possible.* Sitting next to him was Chase *This was the only role Nesbitt would play, and his brief appearance would be the last anyone would hear of the spirit of international cooperation throughout the cold-fusion affair.

xviii B PROLOGUE

Peterson. And to Brophy’s right were Stan Pons and Martin Fleischmann. “We are here today,” Peterson intoned, by way of introduction, ‘“‘to consider the implications ofa scientific experiment.”’ A nonpracticing physician trained at Harvard and Yale, Peterson projected.the avuncular bedside manner ofa family doctor. He was already well known to science reporters. Six years earlier, when University of Utah surgeons had implanted the first artificial heart in Barney B. Clark, Peterson had run the show. Reporters had found Peterson dignified, charming, trustworthy, and very much in control. The New York Times had called him ‘‘a voice of calm in periods of high excitement and repeated crisis.’’* As he introduced the discovery of cold fusion, Peterson’s manner was

so understated that he might have been announcing the hiring of a new professor or unveiling plans for a new medical center. But his even tone was overwhelmed by the flagrant optimism of his message. Over the past several decades, Peterson explained, the U.S. government had spent some $8 billion pursuing research in nuclear fusion. It was considered the energy source that would save humankind: the mechanism that powers the sun and stars, harnessed to provide limitless amounts of electricity. Since shortly after the Second World War, physicists had worked to

induce, tame, and sustain fusion reactions by re-creating the hellish heat

and pressure at the center of the sun in a controlled setting. The conventional wisdom was that sustained nuclear fusion could only be achieved in the laboratory with enough heat—tens of millions of degrees—and extraordinary technological wizardry. Every industrialized nation in the

world, at one time or another, had attempted to create and sustain

nuclear fusion, but progress was slow. Physicists were still decades away from creating a viable fusion reactor. Now, Peterson was saying, these physicists might be out of busines s. The accompanying press release would put it in even more confid ent terms: “Two scientists have successfully created a sustained nuclear fusion reaction at room temperature in a chemistry laboratory at the University of Utah. The breakthrough means the world may someday rely on fusion for a clean, virtually inexhaustible source of energy.” What more was there to say? When Peterson finished his remarks, Stan Pons rose to speak. The forty-six-year-old chemist was wearing a conservative dark blue suit and a polka-dot tie and appeared pale. His hair was cut in a peculiar style that *Coincidentally, the announcement of cold fusion fell on the sixth anniversary of Barney Clark’s death.

PROLOGUE

@ xix

The Boston Globe would later liken to a Buster Brown and the Los Angeles Times to a Julius Caesar. There was something about him—perhaps the hair or his nervous manner or his thick eyeglasses—that made him seem like the archetypal scientific egghead. In his hands, Pons cradled a model of his cold fusion reactor, which looked like—and was—nothing more

than a sophisticated test tube about the size of a highball glass. Pons cleared his throat and looked down at his notes; he spoke quickly and so quietly that his soft southern drawl could barely be heard. He thanked the university for its support and Martin Fleischmann, his codiscoverer, of whom he said, “a finer scientist and person would indeed be hard to find.’ Then he said, ‘““We’ve established a sustained nuclear

reaction by means which are considerably simpler than conventional techniques.” In fact, as Pons explained, their laboratory setup was similar to what might be found in a freshman-level college chemistry course. Then Martin Fleischmann stood, took the reactor from Pons, and

explained that cold nuclear fusion could be sustained “‘indefinitely.” Fleischmann, age sixty-one, was visiting from the University of Southampton, where he had built a name for himself as one of the most distinguished electrochemists in the world. His thinning brown hair was combed over the crown of his head, and he wore dark-rimmed eyeglasses. He stooped slightly and spoke with a clipped and very distinctive Slavic brogue that betrayed his Czechoslovakian birth and English upbringing. Fleischmann seemed tired and ill at ease that afternoon. Of the five men at the table, he may have been the only one who fully grasped the magnitude of their actions. Just that morning he had warned one of the graduate students that their lives would never be the same after they went public; the response, he had said, would be astounding.

To the assembled reporters, Fleischmann described the nature of the experiment that had achieved such remarkable results. In nuclear fusion, he explained, lighter atoms join together—or fuse—to form heavier atoms. In particular, he and Pons had managed to generate the fusion of deuterium atoms, which are a heavy form of hydrogen, by compressing them inside their cold fusion cells. These cells consisted simply of two metal electrodes—one palladium and one platinum—connected to a moderate electric current and submerged in a bath of heavy water (in which hydrogen atoms have been replaced by deuterium atoms). To this was added a dash of lithium. ‘“That,”’ he said, ‘‘is really the guts of the experiment.” They had achieved this stupendous breakthrough, Fleischmann said, on a shoestring budget: ‘We thought this experiment was so stupid that we financed it ourselves.’”” He even showed aslide ofa cold fusion reactor

xx @ PROLOGUE

mounted in a dishpan. This was for lecture purposes only, Fleischmann explained, to demonstrate that they “couldn’t actually pay for very much so it had to be done in a Rubbermaid basin.”’ Even so, they “had burned

up about a hundred thousand dollars” of their own money, which might sound like a lot, but not compared with the tens of billions spent by physicists trying to achieve the same result. For the next thirty minutes, Pons, Fleischmann, Peterson, and Brophy

answered questions from the reporters and scientists. The proceedings were subdued by the standards of a political press conference, orderly and quiet. There was no shouting or interruptions. None of the reporters knew quite what to ask. Nuclear fusion, after all, is a complicated process. What evidence did they have for this sustained fusion reaction? A miniature sun, perhaps, in their basement? In a few short years, Pons and Fleischmann said, they should be able to build a fully operational nuclear fusion reactor that could produce electric power. They said that the oceans of the world would supply fuel for these reactors for millions of years. Although no one explicitly said

that science and technology, once again, had saved the world, that belief

was implicit in their message. In the weeks before the press conference,

Peterson, at least, had envisioned what limitless, clean energy would

mean to a world relying on fossil fuels, a world with serious energy and environmental problems. Not only would it eliminate acid rain and the dangers to the ozone layer but it would negate the United States’ dependency on Middle Eastern oil. There would be no more Three Mile Islands or Chernobyls, no more endless accumulation of hazardous nuclear waste. And, of course, cold fusion represented the potential for making huge sums of money—equal to the wealth of OPEC at least. There were billions and billions of dollars to be made and Nobel Prizes

to be won. Indeed, cold fusion seemed like salvation in more ways than

one. Peterson read a congratulatory letter from Norman Bangerter, the

governor of Utah, in which he said that cold fusion proves , once again,

that “this is the place’’—echoing the words of Bngham Young , who with his Mormon followers had arrived in this promised land in 1847.

As the press conference concluded, however, Peterson may have had

a momentary lapse in confidence. He leaned over to Brophy, who had been trained as a physicist, and asked whether he truly believed that Pons and Fleischmann had discovered cold nuclear fusion. Broph y replied that he had no doubt at all. It was all Peterson needed to reinfo rce his faith. He told reporters that their discovery “ranks right up there with fire, with cultivation of plants, and with electricity.”” Brophy added that Pons

PROLOGUE

@ xxi

and Fleischmann had reproduced their experiment and confirmed their conclusion dozens of times. The data, he said, were “‘overwhelming.”’

By the following morning, cold fusion had made front-page news throughout the world, and it would continue to do so for months. Over the next year the scientific community would spend nearly $100 million trying to confirm the Utah discovery, and cold fusion would become the most controversial, if not the most bizarre, scientific episode in decades. If after the press conference Pons, Fleischmann, Peterson, or Brophy

considered the possibility that they might be dead wrong about the existence of cold fusion, they rarely, if ever, let it show.

7%

bY

sory

Rohe

;



84 Wel pee pee een ce ii? oS ted $5 ieteb eeebiwss + : Ly en rie: het et Bike



wey

“sz,

yrs

d

A

}

Mi

Tw

GOS



USERS a

ot

oT

dt) Di

Ses

ru

eaan

aa}

Pees (Soe rain | ait

oh aman

a

Ary?

macrrig Qube ae So eerieeeAas, soasnee Lat salesietens meet ethyl

in’

Wo!

et

d-Aets

hy. Stave ri kel ci ‘ he

eit

TVR

R

yee

Willies Daeg

(akg, eer ¥

tes

ater

one

bldg wherein

fi

e

a

7

. =

=5

-_

FE

et

a

hat

iwra

,

4



AS

Pet

ak

gS 1

eee &,

-

EVERY

,

OL

2 Pr

¢

nt

Sams

eee

holon, rel

ect

‘= 40

=P

wee

Ap

(eter: Yugi

2a

:

pty dehy

;

a et ie ,

I

Ls Ot

C2 we Paya; it Alenget esmaclie pokes

Paley

sys

: ghee +39 Heats a

Camas tee

© Waa aa

Tae ose inhior “hve aang

peptertnsrtiol sagt cat

:

ee a

Ti / Pere ¥

fr

aN

te

ee Hy Ly oh utesSee oat

Dae

nycoe * 2:

aes ae

sah

a

ee

i

:

,

a



een ove,

Leena et

dey

“asatio>

NN yoedtey anes

eri ware @ enea_ane (erg aes ani

FTI.

Seek atin, Ayelet

ie

1S whe J* aa

CaMER Ete UIP Ticnhir (yes Marke, ng Maman giles (M6 Puss beWa a. Me Ata ee hee mht hep antDa batetzee_ (py

oF5

;

=

2 2

et)

Book J

DELUSION IS THE BETTER PART OF GRANDEUR Anyone with an alertness of mind will encounter during the course of an investigation numerous interesting side issues that might be pursued. It is a physical impossibility to follow up all of these. The majority are not worth following, a few will reward investigation and the occasional one provides the opportunity of a lifetime. How to distinguish the promising clues is the very essence of the art of research. WILLIAM BEVERIDGE, The Art of Scientific Investigation The only way of making clear pea soup is by omitting the peas. A. J. LIEBLING, The Press

-

~2 u

A

4 ’

. 2

:

a. et

Sh

6

Bhi2Nee wetag

oe tt

e

P

i

SaTTI8

7a

|, =

iy

4

ee

wt

2

wee

=

=

é

77

+

*

|

1

MONA

edi yo ek

oF ‘phate? 75s bine bsheee ©

pris! aide om

oo aseta

ai:sae Peit Gk evan

of aa

rabivong: ane ‘ant aisNSoe- > Bn anki; suggest joint effort —Background radiation—what from? how much?

Gajewski mailed a cover letter with the proposal. According to the BYU account, that letter said ‘nothing about declining to review the proposal if the reviewer was doing related work. Indeed, most of the proposals which Dr. Jones is asked by the DOE to review relate closely to his active research on cold nuclear fusion, including muoncatalyzed fusion.’ Gajewski’s responsibility as a funding officer was to have the proposal reviewed by the best-qualified referees, and undoubtedly he felt Jones fit that description. “I think if I did not send it to Steve,’’ Gajewski said later, “I would have been criticized for not using the advice of a person who is perhaps the only person in the United States who is thinking and working along similar lines.”’ That Jones was preparing to launchasimilar line of research may not have mattered, except for a second clause in Gajewski’s cover letter, which stated that the reviewer agreed to “‘use the information contained in the proposal for evaluation purposes only.” This is where the issue turned. Jones certainly had doubts about following up on the review. “Possible conflict of interest,”’ he had told Palmer. He also was immediately aware, or Gajewski made him aware, of Pons and Fleischmann as competitors, or potential collaborators, hence the note in his logbook “‘suggest joint effort.’” He received the Utah proposal and then remarked about the history of his own work. Perhaps Jones himself was simply stunned by the remarkable coincidence. Jones later said he reviewed the proposal because he trusted Gajewski’s judgment, and until he did so he couldn’t know how closely it bore on his work, although Gajewski could and did. Gajewski, more than anyone, knew that Jones was in the market for a new line of research and that the JASON review would not save his program. It seems an unlikely coincidence that both Jones and Rafelski received the proposal from Gajewski in mid-September, just before a

40 8 GARY

TAUBES

muon-catalyzed fusion meeting that Rafelski was hosting in Tucson. All three scientists would be at the Tucson meeting. Thus, they could, and would, discuss what research Jones and Rafelski would pursue next. On September 21, the day after Jones noted his criticisms of the Utah proposal in his logbook, he flew to Tucson for Rafelski’s meeting. That day he composed alist of evidence for the existence of piezonuclear fusion (which he again called mineral-catalyzed fusion).? At the end of the list, Jones wrote, “‘school start-—new neutron counter . . . nearly ready.” On September 22, in Tucson, Gajewski publicly revealed JASON’s pessimistic critique of muon-catalyzed fusion research.? Although Gajewski said that he remained optimistic and that they were not dead yet, Jones wrote succinctly in his lab book, “‘critique ‘Impossible.’ ” At some point during the three days in Tucson, Gajewski, Jones, and Ratfelski discussed the future of the BYU research program. Gajewski recalled that both Rafelski and Jones suggested they move back, as a team, into the piezonuclear fusion work, and Gajewski agreed. If the science was valid, it would be an interesting new approach to fusion. It would not, he said, ‘“‘constitute a continuation of this muon-catalyzed fusion grant.” What was curious about all this was that, when Gajewski sent Jones the Pons-Fleischmann proposal, he was very much aware of Jones’s interest in piezonuclear fusion.* In fact, he later said that when he first received the proposal he was “‘startled”’ at how closely the research resembled Jones’s. He insisted, however, that he could not recall any discussion in

Tucson pertaining to Pons, Fleischmann, or their proposal. “But I do remember,” he said, “that Steve and Jan felt that there is a need to go ahead with this piezonuclear fusion experiment.” The idea that both Rafelski and Jones received the proposal and then refrained from discussing it with Gajewski is difficult to swallow. Jones at first contended that he had not read the proposal immediately, but that argument was not consistent with his logbook, which is to say his September 20 notes on the proposal. Rafelski said that because of the extra administrative burden of preparing a workshop, if he had received the proposal, he would not even have opened it until afterward. “I definitely did not have any intimate knowledge of the Pons and Fleischmann proposal at the time of this workshop,” he said. In any case, while still in Tucson, Jones made a list of things to do. It included a note to get copies of “‘piezonuclear fusion” to Gajewski, apparently referring to his 1986 paper with Clinton Van Siclen.

BAD

SCIENCE

@ 41

Another note read, “‘call Palmer, n/gamma detection, furnace/electro-

chemical—direct students.” Two days later, according to the logbook, Jones and Gajewski shared a ride to the Tucson airport. Once again Gajewski said that it was time to wind down the muon-catalyzed fusion program; his support would last only until November 1989, unless Jones had a breakthrough or went off in a new direction. So Jones once again began to work on cold nuclear fusion, although still by proxy through his students. The assignment went to an undergraduate chemistry major, Eugene Sheely. Sheely had worked for Jones since 1986, mostly on muoncatalyzed fusion, and he’d had his name on a few published journal articles. During the first half of September, Sheely had done related work at Los Alamos. On the seventeenth, he got married and took off to Las Vegas for a honeymoon. When he returned to BYU on the twenty-seventh, one week, at least,

after Jones had read and noted his criticisms of the Pons-Fleischmann proposal, Sheely received a note from Jones requesting that he start doing electrolysis and looking for signs of fusion. Stuart Taylor, a first-year physics graduate student, was assigned simultaneously to try some of Paul Palmer’s electric spark approaches. Jones also arranged with Bart Czirr to use his detector to observe Sheely’s electrolysis cell.* Czirr then began setting up a gamma ray detector, because his latest neutron detector was still out of action. They would have preferred to look for neutrons from deuterium-deutertum fusion because the signal would be more definitive, but they had to use the equipment at hand. Jones also discussed cold fusion with Daniel Decker, chairman of the BYU physics department. Jones showed him Palmer’s two-year-old gamma ray spectrum, with its little bump that Jones thought might be evidence of fusion. “Yeah, well,” Decker said, ‘“‘there might be something there. It is an

awful small bump.” And Jones, as Decker recalled, said, “This is real. Can you help me figure out what is really going on?” Decker then outlined on his blackboard an elementary model of the system. He modeled the electrode in Palmer’s cold fusion cell as a lattice of charged metal ions, stripped of their electrons, then all these electrons flowing freely among the ions. It was as though this sea of electrons was washing around islands of metal ions. The deuterium atoms would enter the metal lattice, give up their electrons to this sea, and then, maybe, fuse. Decker wrote down the equations for the forces between the deute-

42 P GARY

TAUBES

rium nuclei in this sea of electrons. The nuclei would be pulled toward each other by the strong force, the same attractive force that binds the protons and neutrons within each nucleus. Simultaneously, they would be repelled from each other by electromagnetism, because the nuclei have an overall positive charge from their resident protons, and these like charges repel. The strong force, as the name implies, is much stronger than electromagnetism, but it fades quickly with distance. The pivotal question was whether the deuterium nuclei could ever get close enough so that the repulsion of the electromagnetic force would be overpowered by the attraction of the strong force. If that could happen, the two nuclei would become one. According to Newtonian physics, fusion would never occur under these circumstances. But quantum mechanics provides an out, which is to say it assigns every action a probability and, unless the action in question violates some fundamental law, such as conservation of energy,

it would never have a probability of exactly zero. This uncertainty is the fundamental nature of quantum mechanics. In the case of fusion, there is always the possibility, however infinitesimal, that any two deuterium nuclei would “tunnel” under the barrier of their repulsive forces and fuse. This is what Jones had calculated in his paper with Van Siclen in 1986. Tunneling of deuterons is the quantum mechanical equivalent of a snowball’s chance in hell. What Jones hadn’t calculated, however, was the crucial unknown.

What about that sea of free electrons in the metal? Did it somehow manage to cancel the electromagnetic repulsion so that the deuterons could get close enough for the strong force to fuse them? Jones had a graduate student copy the equations off Decker’s blackboard, write them up as a computer program, then plug in the numbers. The result, Decker recalled, was that the probability of the deuterons fusing was, for all intents and purposes, zero. But that was just theory. Steve Jones was an experimentalist. On September 30, Jones sent Gajewski his review of the Pons-Fleischmann proposal. He suggested that it not be approved,’ at least until Pons and Fleischmann could address a number of weak areas, which he enumerated. The last of these, Jones said later, was that Pons and Fleischmann had cited no references. He wrote, ‘““One wonders if a thorough

literature [search] has been done. In particular, publications by C. Van Siclen and S. E. Jones (J. Phys. G. 12 [1986] 213-221) and by B. A. Mamyrin and I. N. Tolstikhin (Developments in Geochemistry 3: Helium Isotopes in Nature, New York: Elsevier, 1984) could be relevant.”

BAD

SCIENCE

@ 43

Jones said he suspected that mentioning his own obscure publication, coauthored with Van Siclen, would be a powerful hint that he had authored this anonymous review. And Fleischmann later remarked that both he and Pons knew from this first go-around that this review had come from Jones. He said of the five reviews, two were positive, two definitively negative, and one equivocal. The last was Jones’s. The piracy scenario that Pons and Fleischmann later constructed was based on their eventual, if not immediate, identification of the Jones review. At least, according to Chase Peterson, this is what they told him

when the growing confrontation came to his attention. ““You see, this proposal of Stan and Martin’s is rejected. [The reviewers] say it’s a worthless thing to promote because we can’t see how it will work. So Stan and Martin think, ‘Do we answer it? Do we tell them more?’ As

they report the story to me they said, “Well, we guess we have to.’ So they give some answers to the questions posed. And the second letter of opinion came back from Jones, still not understanding why it might work, and they finally gave him more data and more rationale, and then the response says, ‘Oh I see, maybe it would be worth doing,’ and that’s what got Stan and Martin very concerned. They felt that until they had explained it in more detail Steve Jones was not aware of the mechanisms by which this could be done. Even though he tried it presumably, and he has these notebooks notarized and so on. But it had failed and he had pretty much abandoned it; then, only after these letters were exchanged, he tried again, and he started to get positive results.”

Since Joey Pons had left in August to get married, the fusion project at the University of Utah had been in hiatus. Apparently Pons was waiting for a response on the proposal and for the money that would come if Gajewski agreed to support him. It was in late October, about the time Pons received the first round of reviews, that he asked Marvin Hawkins

to take over the experiments. Hawkins was old for a graduate student. He was twenty-seven when cold fusion entered his life. He was also the only member of what he would call “‘the local religious persuasion’ in the Pons laboratory. He had taken a two-year mission in Australia, delaying his entry into graduate school. Hawkins had done his thesis research on microelectrodes and had proven an ability to acquire the requisite instruments, build the apparatus, configure experiments, and work with electronics. He was also an accomplished glass blower. To Pons, who was acutely concerned with secrecy, this meant that he could keep his fusion experiments running without recruiting help from the chemistry department support

44 2 GARY

TAUBES

staff, who might talk. Pons even used Hawkins to purchase equipment so he wouldn’t have to go through the usual university channels. And Hawkins wouldn’t talk. He was, at least until March 23 or so,

devoted to Pons. He had grown up on afarm on the plains of western Colorado—“‘chasing cows and raising hay and enjoying the good life” — and found Salt Lake City too “‘big”’ for his tastes and the general population entirely too self-interested. Not only was Pons generous with his time but he was somebody to whom Hawkins could turn when he needed help. As for Fleischmann, Hawkins referred to him as a “god.” Fleischmann’s grandfatherly disposition, he said, had won his heart. Fleischmann later told Hugo Rossi that they gave the cold fusion assignment to Hawkins because “‘we knew we could trust him to do what we told him to do, [and] do it correctly.” Before Hawkins took up cold fusion, he had been feeling claustrophobic at the university, which was a symptom that he had been a graduate student long enough. After four years at Utah, he knew what he wanted. He observed that with a doctorate in chemistry, he could begin work at $40,000 to $50,000 a year. He was earning $10,700 as a graduate student in Pons’s windowless basement laboratory. On that salary he had to support a wife and three children, which was not possible even in Utah. He was surviving with guaranteed student loans. “I was tired of being low on the totem poles,” he later recalled. ‘I was tired of running. I wanted to finish and go on.’’ Another reason Pons recruited Hawkins out of his dozen or so postdoctoral and graduate students was that Hawkins had just completed his thesis research, so all that remained was the writing. When Pons requested that Hawkins fabricate cells and take data on the fusion project, Hawkins figured the work might postpone his entry into the real world by as long as six months. He also considered the fact that the research could lead to a share in a Nobel Prize: “I'd have been an idiot to have said no.’’ Once Hawkins took up cold fusion, he began to think that maybe it was his destiny. When Hawkins took over the experiments in late October, his mentor’s concern with nuclear radiation was still limited. Said Hawkins, “If

we're getting zapped with heavy doses of radiation, we want to know about it.” Once the reviews on his proposal came back, or at least about that time, Pons investigated obtaining some radiation detection equipment for his laboratory. Pons called Robert Hoffman, a health physicist on campus, and asked what kinds of instruments he would need to measure neutrons, protons, or gamma rays.’ The question suggests that Pons had not given much, if any, thought to the actual physics of nuclear fusion before that time. It

BAD

SCIENCE

@ 45

also suggests that Pons was responding to his reviewers, who would surely have suggested that the standard evidence for the existence of a nuclear reaction is nuclear radiation—in this case, neutrons or gamma rays. Hawkins, meanwhile, had spent November designing and running cells by trial and error. Pons had ordered and received four palladium rods to use as electrodes, and Hawkins was trying to create the perfect fusion cell for them. It would take him a week to put a practice cell together. Then, as Hawkins put it, he’d “‘try it, crash it, retrofit, change

and redrill and rework, take maybe another four or five days,.and try something different.” Hawkins said that until he began his trial and error program, they had been running only one fusion cell in the lab. This apparently was the one that Fleischmann would later refer to as their “‘monster cell.”’ The monster’s electrode was a sheet of palladium, eight centimeters square by two millimeters thick, which was the highest gauge commercially available. It was bent in a U-shape and placed inalarge glass vessel with a platinum electrode in the center. For a variety of reasons, Pons had not been happy with the monster, although they apparently kept it charged. When Hawkins joined up, Pons explained to him that he wanted a cell that would allow him to monitor and control carefully the temperature emitted by the cell, while distributing a uniform current around the electrode. The monster apparently failed to fulfill both of these requirements. Hawkins said his first cells were beset by a variety of problems, the most stubborn of which was a tendency of the palladium electrode to break or short out as it absorbed deuterium and expanded. Hawkins beat this by building a cage of four glass rods around the palladium rod, then winding the platinum anode around the glass cage. This entire assembly fit inside a double-walled glass vessel, known as a dewar flask, which was sealed off at the top bya plastic stopper. (Hawkins’s account implies that the infamous fusion cell, which later would be imbued with nearly magical powers, had been designed by Hawkins himself, a graduate student who had worked by trial and error with little guidance from either Pons or Fleischmann.)

Forty-five miles down the road in Provo, Steve Jones was neither waiting to throw himself into cold fusion, as Pons and Fleischmann would later surmise, nor deviously plucking the two chemists for details. He was simply letting Eugene Sheely, his student, work on it while Bart Czirr and Gary Jensen helped with the detectors.

46 # GARY

TAUBES

Sheely had read up on electrolysis and asked James Thorne, his adviser in the chemistry department, about it. He said he found it not particularly difficult to learn. He spent a few days procuring the necessary equipment and the heavy water. He used sodium sulfate for his electrolyte, because he had it in the lab. He assumed any salt would do, even table salt, as long

as it didn’t corrode the electrodes. And Thorne provided a power supply and suggested he use stainless steel electrodes. When those turned out to rust quickly in the electrolyte, Sheely switched to nickel and eventually palladium because, he said, Jones suggested several times that he try palladium. Why palladium? Jones insisted it had nothing to do with what he might have read in Pons and Fleischmann’s proposal. According to Jones, Thorne suggested palladium as the natural choice because it absorbed so much hydrogen. At any rate, Jones included it in the list he had written down back on April 7, 1986, as a metal that “‘dissolves much hydrogen.” That was one way of looking at it. Another was that after reading the Utah proposal Jones did indeed discuss this issue with a chemist, Thorne, and then assigned a chemist to the electrolysis instead of one of his physics graduate students. This train of events was consistent with the pirating scenario that Pons and Fleischmann later constructed. Jones, the physicist, seemed to be pursuing cold fusion from an electrochemist’s perspective. Impetus, to use Jones’s word, appeared to be present.® Either way, Sheely eventually got his cells running but spent most of his time waiting for Czirr and Jensen to get the gamma ray detector running as well. The device seemed to be in a constant state of disrepair. Jones came down to see the work a couple of times, and he had a group meeting once a week, during which he always asked Sheely for a progress report, but he seemed to be preoccupied with the problem of getting money for muon-catalyzed fusion. Even after the Tucson meeting, according to the record in Jones’s lab books, both Gajewski and Jones were still determined to find money to save the program. On November 8 Jones spoke at length to Gajewski about muoncatalyzed fusion. Jones wrote in his notes that Gajewski had looked into DOE’s conventional fusion programs as a possible home for it, but he had “little hope.” And DOE’s Atomic and Molecular Physics Division, which would also be a possibility, was “clogged” with projects and had “little turn-over.”” Finally Jones and Gajewski discussed what is known disparagingly in scientific circles as earmarking, appealing directly to Congress for support, thus bypassing entirely the DOE bureaucracy, peer review, and all

BAD

SCIENCE

& 47

the attendant frustrations. Jones sketched out a possible plan of attack: Define the best congressional committee, approach the committee, choose one of his local representatives to go through, and so on. With luck and perseverance, he could end up with a bill in Congress that said that so many dollars had to be spent by the Basic Energy Sciences Division of DOE on muon-catalyzed fusion. Still, earmarking was considered a last-ditch contingency. From November 9 onward, Jones began to attend personally to piezonuclear fusion. On November 16 he wrote in his logbook, “‘pons response,”’ among alist of things to do, and he checked it off. This was his first mention of Pons in his book since September 29. Apparently, Jones had now sent in his review of Pons’s revised proposal. Jones said later that this was the review in which he recommended that the PonsFleischmann proposal be approved. On November 28 Jones spoke to Rafelski, who apparently told him that Pons ‘‘needs nuclear physics help.” Jones noted that “Jan wrote of their ignorance, rejected proposal.” The next entry in the logbook is December 1, and it finally concerns Sheely’s electrolysis work. Sheely had now given the results to Jones in an organized form. On top of the page, Jones wrote, “Pd + H,O + D,O runs show small excess of neutron rate vs background.” Sheely later translated this as “exciting but definitely real close to background. Nothing anyone thought was publishable, we needed more data, but it was enough that it was more than worth continuing to test it.” Then come several pages of calculations working out the signal and the background from Sheely’s data. All the runs seem to have been recent and Czirr’s neutron detector, it seems, was finally running. There are

four pages of these calculations, the last of which is dated December 5. On December 9 Jones was in Washington, meeting with Gajewski and Rafelski. Once again, Gajewski ran down possible sources of money for muon-catalyzed fusion, and Jones actually drew in his logbook a schematic diagram of the DOE hierarchy. Evidently the directors of the conventional fusion programs had declined to offer support for Jones’s work. But the division head at Atomic and Molecular Physics would accept a short preproposal. Jones wrote, ‘Congressional route a possibility.” Also possible were the National Science Foundation and EPRI, the privately owned Electric Power Research Institute. Then the discussion turned to cold fusion. Jones described Sheely’s results. Then comes this note, at the top of page 59:

48 § GARY

12-9-88

TAUBES

stimulating nuclear fusion by means of flow of hydrogen isotopes in metal lattice

/ Jan: 1—Patent: Jones, Rafelski, Palmer

Rafelski apparently suggested that they consider patenting the discovery of “stimulating nuclear fusion by means of flow of hydrogen isotopes in metal lattice” and that the names on the patent should be Jones, Rafelski, and Palmer.’ This is a remarkable suggestion, especially when made by two reviewers of a proposal on the identical subject that, if nothing else, had provided an “‘impetus”’ to pursue their work. The names of Pons and Fleischmann are conspicuously absent here. As none of the principals professes to remember details of the conversation, this is the only record that exists. Gajewski then told Jones that he could also submit a proposal on cold nuclear fusion, independent of the Pons-Fleischmann proposal. On the flight back to Salt Lake City, Jones composed a draft of what was either his piezonuclear fusion proposal or a paper claiming its discovery. Jones does not remember which, he only remembers writing it. The title is “First Demonstration of Cold Piezonuclear Fusion.” The six pages begin with Jones’s work in muon-catalyzed fusion and describe how that led them to believe that there must be other ways to induce hydrogen atoms to fuse. After discussing the geophysical evidence for such fusion, the draft sketches a quasi-theory that would allow hydrogen nuclei to get close enough to fuse in a metal lattice provided two hydrogen ions could inhabit the same site. This, Jones wrote, would be enhanced “when the ions are rapidly migrating in the host metal, jumping between interstices.” Thus, with what could be called either vigorous hand waving or a miracle, “cold piezonuclear fusion in metals (one could call the process metal-catalyzed fusion) appears feasible at a very slow rate.”’ These considerations, Jones continued,

‘‘have motivated a series of

experiments at Brigham Young University to search for metal catalyzed fusion.” He then described Sheely’s recent electrolysis experiments and wrote that they had seen both gamma rays and neutrons with a “specially-designed neutron counter [that] was built for this experiment.” He ended with his characteristic optimism: In conclusion we have demonstrated for the first time that nuclear fusion occurs when hydrogen and deuterium are electrolytically loaded into a metallic foil. This remarkable process obviates the need for elaborate

BAD

SCIENCE

&@ 49

machinery to generate and contain either plasmas or muons to induce fusion. We are now exploring means to enhance the fusion yield of this

new process.

At this time Jones was still considering two strategies. At some point after receiving the first version of the Pons-Fleischmann proposal, he suggested that Gajewski inform Pons outright of the existence of Jones and his research. Jones said he may have made this suggestion immediately upon receiving the proposal, as early as September 20, which his logbook bears out. Yet Gajewski didn’t act on it until December 16, after getting the second round of reviews back, after Jones had time to get the results from Sheely’s electrolysis experiment and discover that there might be something to this phenomena, and after Jones, Rafelski, and Gajewski met in Washington to discuss the future of cold fusion. Jones later admitted that he considered keeping his research secret and going public first. This gambit appears several times in his lab books between December and March. ‘“‘There’s this little closet under our stairs, carpeted,”’ Jones said, ‘“‘and I’d go in there and I would think, and I would pray, and I would think, What do I do? How doI handle it? I’ve

read their proposal. I’ve been working on this, and I don’t deny there could have been some impetus. And the solution I came up with was tell them and let them use our detector.” Gajewski apparently favored a similar strategy. The way he saw it, Jones had the physics expertise and the radiation detection equipment. Pons and Fleischmann had the chemistry expertise. They were attacking the same problem from distinct but mutually supportive directions. “Pons was working on heat,’’ Gajewski said, “‘and Steve was working on neutrons. And I would have been delighted at the time to fund both, one to continue working on neutrons, the other to continue working on heat, and possibly collaborating and checking on each other.” It made perfect sense, except that at this time neither had any data worth mentioning. Still, according to Jones, Gajewski even took into account ‘‘that a cooperative effort from the nearby schools could provide an efficient use of taxpayers’ money if things worked out and their proposal was approved.” Bart Czirr wanted to sell Pons and Fleischmann one of his detectors, and he understood, as did Paul Palmer, that Gajew-

ski was telling Pons and Fleischmann that they wouldn’t get funded without a decent neutron detector. ‘“‘Gajewski told them to get down here and learn how to measure neutrons,” said Palmer. ““They couldn’t

get funded in nuclear physics if they didn’t know how to measure neutrons. Gajewski, in my opinion, was doing the best job a civil servant

50 8 GARY

TAUBES

could do. He said, ‘Why don’t you stupid guys cooperate?’ [To us, he said,] “You need electrochemical expertise. You’re doing electrochemistry, and you’re doing it ignorantly and stupidly. Why don’t you get some electrochemists?? And he told them, ‘You guys don’t know anything about nuclear physics. Why don’t you get some nucleat physics help? Some detectors. Why don’t you work together?’ ”’ Finally, on December

16, comes this note in Jones’s log:

R. Gajewski—spoke to Pons, who had my rebuttal & Mamyrin paper —benefit to science aided by interaction but no pressure to work w/ Jones —get together & measure each other —spoke of ’86 report to him (Jones—mcpf)'°

And later that day, Pons and Jones apparently spoke for the first time: Pons: several weeks to load p-d to start expt. interested in our neutron detector! seeks help

According to Jones, he then sent Pons the requested information about Czirr’s neutron spectrometer, and Czirr sent along one of the proposals he had drafted for the Department of Energy. Czirr recalled that he had carefully chosen this material so that Pons and Fleischmann could see what he had to offer but not use the information to build a detector themselves. Although from Jones’s notes Pons seems to have been amenable to a working relationship, that was not the case, even this early in the game. If Pons already believed he was sitting on the discovery ofa lifetime, how was he expected to take these two phone conversations? Pons would later say these calls were the first time he ever heard of Steve Jones, and the BYU physicist was immediately suggesting they work together. Gajewski may have been the first funding agent to whom Pons showed his proposal who could have been considered a stranger of any sort. Up until Gajewski, Pons had been reluctant to tell his friends about cold fusion. Even when he told Richard Bernstein of UCLA about fusion in September 1987, he swore Bernstein to secrecy. Now Gajewski called him and suggested that he collaborate with one of his reviewers. Gajewski also told him of the report that Jones had filed in 1986 on piezonuclear fusion, which is to say cold fusion. Gajewski “spoke of ’86 report to him” is how Jones recorded it in his notebook. So Gajewski, whether he realized it or not, was now telling

BAD

SCIENCE

@ 51

Pons that Steve Jones, his reviewer, had been doing the very same work since the spring of 1986. (Quite a coincidence! So Jones, the reviewer, was first.) Stan Pons could not have been pleased. He was alone; Fleischmann would be in England for quite a while. And he was being robbed. Although Gajewski suggested that Pons collaborate with Jones, it must have appeared to Pons as though Jones and Gajewski had decided that they were moving in on his territory. It seemed like a fait accompli, which in a sense it was. It was around this time that Pons met Eugene Loh, the chairman of the Utah physics department, at a campus Christmas party. Pons asked him where to obtain a neutron detector and actually described his experiment. Loh later said the concept seemed so ludicrous that he assumed Pons was joking.’ As for the neutron detector, Loh directed him to the Los Alamos National Laboratory. ‘Pick up a phone,” he said. “The people at Los Alamos will be glad to lend you a detector.” Instead, Pons seems to have called back Robert Hoffman, the health

physicist, who lent him a small portable detector. This was a dosimeter, which would effectively do nothing more than alert Pons to the possibility that he was getting zapped. “If that thing ever went off,’”’ Marvin Hawkins said, ‘“we knew it was time to shuffle on out.’’ Of course, if his

electrochemical cells were inducing any appreciable level of fusion, as Pons believed they were, the dosimeter should have been more than sensitive enough to pick it up. It registered nothing. By the Christmas holidays, Hawkins had finally constructed cells that he considered ‘“‘reasonable.” Both he and Pons worked through the vacation. As Hawkins recalled it, they were in the lab “all day every day through this period of time.” Pons and Hawkins were now ignoring the issue of nuclear radiation and concentrating on the first of endless measurements of the heat generation from the cell, a procedure known as calorimetry. The purpose of these measurements was to compare the energy going into the cell with the energy leaving it. This required, among other things, constant monitoring of the current and voltage feeding the electrodes and the temperature of the electrolyte in the fusion cell. If carried out correctly, the procedure would show whether the energy emitted by the cell was greater than the energy it took to run it. If this were the case, as Pons, Fleischmann, and Hawkins would come to believe, then the

excess energy could be attributed to some energy generation in the cell, perhaps some unknown species of radiation-free nuclear fusion. Richard Steiner, associate chairman of the chemistry department, said

52 EP GARY

TAUBES

Pons had talked to him about calorimetry and how difficult it is to do it right. “He talked to me,” Steiner said, ‘about having read this book and

that book and looked at this method versus that method, and worrying about how to do it.” Fleischmann, however, later told Dave Williams,

an electrochemist at the Harwell laboratory, that he had. a foolproof method for doing calorimetry. As it would turn out, there is nothing simple or foolproof about calorimetry.

CHAPTER JANUARY

AND

4 FEBRUARY

1989

In the organized competition to contribute to man’s scientific knowledge,

the race is to the swift, to him who

gets

there first with his contribution in hand.

ROBERT MERTON, The Sociology of Science

y early January 1989, Steve Jones was finally dedicating the better part of his time to piezonuclear fusion. Now his own words, such as “vigorously pursuing” or “running as fast’? as he can, could almost be deemed appropriate. Bart Czirr also had his neutron detector running, albeit not well. Gary Jensen, Czirr’s partner, said they switched from the gamma ray detector to the neutron detector only because the background noise picked up by the gamma ray detector made it impossible to observe anything. Detecting neutrons, however, was a perverse endeavor in its own right: on the one hand, neutrons have no electromagnetic charge, which makes them nearly invisible to most detector schemes. On the other, the world is full of naturally occurring neutrons from cosmic rays, known as background radiation. In order to prove the existence of cold fusion, Jones’s experiment had to differentiate between the natural abundance of neutrons in the laboratory without any fusion cells running—the background—and the abundance of neutrons in the lab when the cells were 53

54 © GARY

TAUBES

running—the signal. If the signal was greater than the background, then maybe they had something. Czirr’s neutron detector was a clever modification of a standard design for such devices. It used a material called a lithium glass scintillator, which, in this case, consisted of plates four inches square and a few millimeters thick. If a neutron passed through the glass, it stood a good chance of being captured by an atom of lithium. The lithium would then emit an alpha particle—a helium nucleus—which would have a positive charge, and this would create a pulse of light as it passed through the glass. This light pulse would be detected and amplified and would eventually register as a neutron. Since 1986 Czirr had been modifying the device so that it could also measure the energy of these neutrons, which could be valuable information. Deuterium-deuterium fusion, for example, always results in a neu-

tron of exactly 2.45 million electron volts (MeV). If Czirr’s detector could measure the energy of a neutron, this would simplify the chore of proving whether the neutron came from fusion or some background source. Czirr said the resolution of his detector still fell considerably beneath his expectations: a 2.45 MeV neutron, for example, might show up anywhere between 1.00 MeV and 5.00 MeV. Czirr would have preferred to spend his time working on this resolution problem, but instead he was continually struggling to keep the neutron detector and its attending electronics operating. ““We had old borrowed equipment,” Czirr said. “It’s a bunch of junk really. Half of It, anyway. So it was hard to keep it operating and not running out. So we had our eyes focused on the detection of neutrons and the detector itself. Is it working or is it not?” Jones, his various students and colleagues were still working at what constitutes a leisurely pace in the research community. When Jones eventually got around to publishing, his data consisted of fourteen experimental “‘runs,”’ all between December 31 and the first week ofMarch, and lasting from eight to twenty hours each. This works out to maybe two runs per week. So they were running cells every second or third day

and regularly changing the electrodes and the mixture of the electrolyte, hoping for some combination that would produce copious neutrons. Now and then, they would leave Czirr’s detector on at night without any cells running to measure the neutron background. Jones assembled the data almost daily. “He was watching it all the time,”’ said Czirr, “seeing if there was any hint of a signal sticking its nose up above the background.” They had intimations on occasion, but nothing that a physicist would

BAD

SCIENCE

@ 55

consider compelling. On January 5, for instance, Jones wrote that they had observed a signal using a titanium electrode and sodium sulfate for the electrolyte: Ti + d + n + Na,So,—very high n° rate Fusion neutrons are highly suspected!

A week later, Jim Thorne, the chemist, suggested that they try adding metal ions to the electrolyte. This procedure is known as poisoning, which would help the palladium absorb hydrogen and deuterium and might increase the fusion effect, if there was one. During these first weeks of January, Jones was also writing up his cold fusion proposal, and Jan Rafelski seems to have been spurring him to submit it quickly. On January 10, for instance, Jones wrote in his lab book that he spoke with Rafelski, who suggested an aggressive strategy: Jan: Proposal—submit urgently—on basis of own findings —valid project

That the project was valid seems reasonable, but it is hard to understand what Rafelski and Jones considered the pertinent findings to have been on January 10. They still had no neutron signal worth taking seriously. They did have Paul Palmer’s geophysical data, and by several accounts it was those data that convinced Jones that he had discovered cold fusion and thus should proceed with the proposal. Until January, Palmer’s evidence had depended on excessive amounts of helium 3 that had been found here and there but that could be explained away without invoking cold nuclear fusion. Palmer always thought that if they were to discover excessive amounts of tritium rising from the earth, that would be much more definitive. Tritium is a radio-

active isotope of hydrogen, and, unlike helium 3, it has a very short life span. The half-life of tritium, which is to say the time it takes half of any given sample to decay radioactively, is only a dozen years. If the earth happened to be oozing tritium in excessive quantities, professional geologists could not explain it away as primordial residue, as they could with helium 3. As Palmer recalled it, one of the graduate students, Stuart Taylor,

located at the library a government publication by H. G. Ostlund and A. S. Mason

of the University of Miami,

Florida, called Atmospheric

Tritium 1968-1984. Among the voluminous data recorded by Ostlund

56 # GARY

TAUBES

and Mason were the records of a monitoring station on the slopes of Hawaii’s Mauna Loa volcano, which had recorded atmospheric tritium concentrations from mid-1971 through the end of 1977. Palmer discovered in this data a near doubling of the tritium level during February and March of 1972. This happened to coincide with an eruption of the Mauna Ulu volcano, just twenty-five miles upwind. Later in the year Ulu erupted again, and again the tritium level increased, although not as dramatically. To Jones and Palmer, the Mauna Ulu tritium seemed like a godsend, but the data were not quite so cut and dried. The picture was complicated, for instance, by atmospheric H-bomb tests, which had been blow-

ing copious amounts of tritium into the atmosphere for forty years. And nuclear submarines had been flushing their fair share into the oceans. Nuclear reactor facilities and nuclear weapons facilities—like that at Savannah

River,

South

Carolina,

which

made

tritium

for

the

H-bombs—had also been known occasionally to leak tritium into the atmosphere or groundwater. When this happens, neither our government nor those of the other nuclear nations are likely to advertise it, so it was doubtful that these technological sources could be ruled out as the source of the Mauna Ulu tritium. Man-made tritium was migrating through various ecosystems in startling quantities. The Soviets had exploded a few hydrogen bombs underground just five months before Ulu’s first eruption, and the United States did the same a few weeks later. Nonetheless, Ostlund and Mason seem to have

believed that these explosions could not account for the Mauna Ulu tritium, which gave Jones and Palmer more faith than ever in cold fusion. Jones later said that they checked with the U.S. Navy, which reported, of course, that no nuclear submarines were leaking tritium during that time. And Jones believed them, so that was that. “‘I’ll tell you,” said

Stuart Taylor, ‘““Steve was excited when he saw that tritium data.” As Jones would later put it in the first draft of his cold fusion paper, “We conclude that this volcanic eruption freed tritium produced by geological nuclear reactions.””!

On the weekend of January 28, Dave Williams stopped off in Utah on his way back from New Zealand. Stan Pons told him that they were working on something ‘‘very hush-hush.”’ Williams remembered that back in November Martin Fleischmann had told him “that he had this really interesting thing which was of potentially mind-blowing importances; While he was at Pons’s lab, Williams noticed that Pons and Marvin

BAD

SCIENCE

@ 57

Hawkins were busy unpacking ascintillation counter, a device for measuring tritium levels. Hawkins had already taken samples of the electrolyte from the fusion cells and had them checked for tritium at a radiobiology lab and at another chemistry lab. The results had seemed promising, so Pons had ordered his own scintillation counter. Pons and Fleischmann later reported that they had observed one hundred disintegrations of tritium per minute per milliliter of heavy water, which means that one hundred atoms of tritium were decaying each minute from a one-milliliter sample. This corresponds to a total of 1 billion tritium atoms in each milliliter, a misleading number because it sounds impressively large. Heavy water, because of how it is produced, is likely to contain two or three or five times as much tritium as this naturally.2 Pons and Fleischmann seemed unaware of this fact and considered their tritium data more evidence for fusion. After all, tritium can be created only in a fusion reaction. They were worried, however, because this number was still a billion times less than what they were expecting, considering the amount of heat that the fusion in their cells appeared to be generating. Until the tritium counter arrived, Pons had managed to keep his fusion project successfully hidden from his other researchers. They knew something was going on, but they didn’t know what. Afterward, they started hearing the word fusion.

By January 27, Jones was convinced that he had sufficient evidence of cold fusion to commit himself, and he debated a range of strategies. He could send Ryszard Gajewski his proposal, leaving Pons and Fleischmann to their own fate. He could write up his paper and appropriate the discovery. Or he could take the more charitable route and include Pons and Fleischmann on a joint paper or a joint proposal or both. If he chose the aggressive strategy, as he noted on January 27, it would have to be with a quick and dirty publication of his quick and dirty experiments. Jones realized that this strategy involved taking a stand on his meager neutron data while admitting that he had little understanding of what he was seeing or why. He seemed aware that the geological evidence was still a flimsy foundation on which to build a theory of fusion.’ (“Claim open Ql[uestions]s’’ he writes.) In late January, Jones sent a draft of his own cold fusion proposal to Rafelski for his comments, and then to Gajewski. On January 30 Jones noted in his logbook that Rafelski suggested he mention in the proposal his “‘early and continued interest.’’ The intent is cryptic. One likely meaning is that Rafelski was less sanguine than Jones about the brewing

58 # GARY

TAUBES

conflict of interest. Once Jones had received the Pons-Fleischmann proposal, it could be argued that he was, in effect, damned. Any action on his part after that involving cold nuclear fusion could spark Pons and Fleischmann to accuse him of scientific thievery. He would have to scramble to defend himself. Gajewski later insisted vehemently that there was no conflict; the two proposals were not on the same subject. ‘‘Pons and Fleischmann were doing heat,” he said, “and Jones was doing neutrons.’’ But this point was questionable as long as both proposals involved electrolyzing heavy water with palladium electrodes to induce nuclear fusion. Gajewski kicked the first draft of the proposal back to Jones, directing him, among other things, to get an electrochemist to help with the chemistry. Gajewski apparently hoped that Jones would go back to Pons and Fleischmann, looking for a collaboration. Instead Jones enlisted the support of Douglas Bennion, a prominent electrochemist on campus. As Jones recalled, Gajewski also began urging him to go public with the data, saying, “It will be better if you publish first and then it will go more easily.” Both Czirr and Palmer also picked up this impression along the way, although it’s likely they got it from Jones. Palmer even came to believe that Gajewski said he could approve their proposal without an external review if they would publish a paper.* Gajewski would later contradict this. In his version of events, Jones asked him if he should publish, and he merely suggested that ifJones was ready to publish he should. ‘‘Obviously it would help,” Gajewski said later, “but at the same time, that was not my motivation. My motivation was that if he had something, there is no point sitting on the data because a guy in Singapore may be working on the same thing. That’s elementary in science.” This seems a bit disingenuous. Gajewski surely knew of a guy or two in Salt Lake City working on the same thing. Either way, Jones took this suggestion as a mandate to announce his results. Jones had already been invited to speak on muon-catalyzed fusion at an American Physical Society meeting in early May—the meeting of the year for the physics community. It is usually held near enough to Washington so that the scientists can visit their funding agents, interested politicians and bureaucrats, and suitably impress them. On February 2, one day before the deadline for submitting abstracts for the meeting, Jones sent his by overnight mail. He briefly described the BYU work on muon-catalyzed fusion, then added Just enough on cold fusion to constitute an announcement of discovery: “‘We have also accumulated considerable evidence for a new form of cold nuclear fusion

BAD

SCIENCE

#59

which occurs when hydrogen isotopes are loaded into materials, notably crystalline solids (without muons). Implications of these findings on geophysics and fusion research will be considered.” Employing an abstract to claim a discovery has a long and illustrious history in science. For instance, Edmond

Halley, for whom the comet is named, urged Isaac

Newton to do so in order to secure an invention to himself “‘till such time as he would be at leisure to publish it.” Jones later claimed that this was not what he had in mind. But when he was told that it was this abstract that partially led Pons to believe he was not dealing in goodfaith, Jones responded with an ambiguity worthy of a candidate for public office. ““You’ve seen the abstract,”’ he replied. “It was nothing. Yeah, ‘We’ve seen evidence for this. . . .” I think it was kind of funny.” About this time Jones accepted an invitation to lecture on muoncatalyzed fusion at Columbia University on March 30. Amiya Sen, a plasma physicist at Columbia who had invited Jones, recalled that when he called him the conversation had a curious twist: “‘[Jones] said, “By the way, I’m working onalittle something else, and I’d like to talk on that also.’ I said, ‘Oh fine.’ Then he said, I think, ‘There is a competitor hot

on my tail.’ ”’ On February 3, Lee Phillips got a call from Norm Brown, head of Technology Transfer at the University of Utah. Phillips was Brown’s counterpart at BYU, and the two had acollegial relationship. Brown said he needed to talk to Phillips “about the possibility of one of your professors’ pirating one of our professors’ stuff. What do you know about it?”’ Brown had been talking to Pons about the chemist’s suspicions and assumed he could look into the matter with a simple phone call. He was

still unaware that nothing in cold fusion would be simple. Phillips told Brown that he knew little about the matter, and then asked Carol Hardman, who was responsible for overseeing BYU’s out-

side research contracts. As it happened, the day before, Jones had informed her that he had submitted a proposal to the Department of Energy but that his funding agent had asked him to broaden the scope and resubmit it. Jones also mentioned that some chemists at the University of Utah were doing similar work, which seemed innocent enough. Hardman assumed that Jones was debating whether to include the Utah chemists in the proposal. She relayed to Phillips this misconception of the issue, which he relayed to Brown: Jones had been working on cold fusion for a long time, he said, and the BYU administration had been

keeping an eye on it.

60 8 GARY

TAUBES

This seemed to satisfy Brown. He said he’d call back if they “needed to get together on it.” Hardman thencalled Jones and told him that Norm Brown from Utah had been asking about the provenance of his proposal. Jones, concerned that Pons was spreading unjustified accusations, called Gajewski, who said he’d look into it. In the meantime, on February 8, Jones met with a BYU patent review committee headed by Lee Phillips, regarding a “device to Produce Controlled Nuclear Fusion.” Jones and Palmer presented this reactor-tobe, according to the minutes of the meeting, as though there was no prior work in the field, other than that by a “Russian author.’’ The minutes continue: ‘‘Possible applications would be plastic explosive detection and a small energy source, i.e., for spacecraft. It is anticipated that a neutron source is still 3 to 5 years away. Work is in the early stages. No device is fully developed.” The meeting concluded with Phillips, Jones, and company agreeing that Jones should inform DOE that he will retain patent rights on the invention—as required by DOE—at which point they could take up to two years to decide whether to actually proceed with the patent applications: “The technology has unlimited potential, but it will still take many years to produce a marketable product. The technology is so technical that patent costs will be high, possibly $15,000 to $25,000. More information needs to be obtained before a decision can be made.”’ That Jones, or somebody in this meeting, thought the technology had unlimited potential may be taken as indicative of how premature Jones’s data still were. In any case, the patent review committee then decided that the only remaining concern involved waiting to apply for the patents themselves: “That is, should someone file a patent application claiming the matter which you may have priority to, it would be more difficult and expensive to fight the application than it would have been to file ours fitstene On February 10, Gajewski finally reached Pons about the Norm Brown phone call of a week earlier. Pons then talked to Jones directly, which

appears to have been the first time the two had spoken since December.

Pons apologized for Brown’s intervention. He said they had “‘an overeager lawyer” at the university who had overstepped his bounds. Then either Pons told Jones or Gajewski told Pons—Jones couldn’t remember which, and his notes were vague—that Gajewski was going to hold on to Pons’s funding until the conflict was sorted out. “I think,”’ said Jones, “that Pons didn’t like that.”

BAD

SCIENCE

@ 61

In any case, Jones made no mention in his logbook of any talk of collaboration. On February 14, Fleischmann visited the Harwell laboratory, where he was a consultant. Harwell is the largest applied research laboratory in Bnitain, and is often compared with the Los Alamos Laboratory in the States. It is one of the principal laboratories of the British Atomic Energy Authority and has three working fission reactors. Fleischmann’s visit was specifically to procure the kind of nuclear physics help that Jones was offering. Fleischmann had a private conversation with the research director of the lab, Ron Bullough. Like Fleischmann, Bullough was a fellow of the Royal Society, but, unlike Fleischmann, he was a professional physicist. He recalled that Fleischmann was genuinely concerned about the security aspects of his cold fusion invention. He told Bullough about the meltdown and added that they had seen nuclear signatures—in particular, tritium and neutrons—from other working cells. Fleischmann said that they had to be very careful how they handled it. ‘“He felt he might be wrong,” remembered Bullough. “There was an undercurrent of “This is incredible and I have to be careful.’ ”’ Bullough had trouble believing what he was hearing, although he took it seriously enough. “Given the man’s background and his talent,”’ Bullough said, ‘‘to say he was seeing nuclear signatures out of an electrolytic cell was, to put it mildly, astounding.” Fleischmann gave Bullough a draft of the Department of Energy proposal and asked him to discuss it confidentially with his physics colleagues. He then asked if he could borrow a neutron detector, and Bullough agreed. The Harwell staff took a detector, known as a Bonner sphere, off their production lines and sold it to Fleischmann for a few thousand pounds. (“I thought we had lent it to him,” Bullough said later. “T am slightly embarrassed about that.”’) Finally, Fleischmann suggested that he and Pons might need help in confirming their results, and Bullough readily agreed to that as well. Although it was left unsaid, Bullough believed that Fleischmann had also confided in him to give his country enough advance notice to reap some of the benefits of cold fusion. Fleischmann was anxious that the Americans not monopolize the technology. He wanted Harwell, and England, to be prepared to move fast should he and Pons have to go public. Over the next few days, Bullough studied Fleischmann’s proposal and discussed it with the head of theoretical physics. “I have to say,” Bullough recalled, “there was complete incredulity.”

62 BE GARY

TAUBES

Fleischmann arrived in Utah with the Bonner sphere a week later.® The Bonner sphere was a crude instrument designed for health physics purposes only. It was covered with a layer of plastic and was about the size of a human head, thus simulating the surface area of a human in proximity to a radiation source. If Pons and Fleischmann’s cold fusion cells had been producing trillions of neutrons each second, as they should have if they were generating watts of fusion energy, then the device should have easily been sensitive enough to detect it. It registered nothing. On Thursday, February 23, Fleischmann and Pons finally took Steve Jones up on his two-month-old invitation to visit Bngham Young University and look at their neutron detectors. The two chemists got a tour of the BYU experiment. Whatever Pons and Fleischmann had in mind with this visit, it came off as an acute exercise in miscommunication.

Jones and his colleagues had still been unable to detect enough neutrons to prove that they had generated fusion in their cells. When they went public one month later, it was with a reported signal of two neutrons per hour, which represents roughly one trillionth of a watt of power. Pons and Fleischmann, of course, believed that they had generated enough nuclear fusion to melt a cube of palladium. They believed they were producing untold watts of power. Ifso, they should have been seeing trillions of neutrons per second, but, according to their Bonner sphere, they were not. By the end of the day, as Dan Decker, head of the BYU physics department, recalled, “it was obvious that we had no idea really what they had done, and they didn’t quite realize the minusculeness of what we had found.” The visit left the BYU group with a smattering of impressions. Decker, Czirr, and Palmer all remarked on the fact that Pons had said

remarkably little the whole day. Czirr added that he found Pons quite likable. “I have something of a prejudice for southerners,” said Czirr, who was born in Oklahoma. ‘“‘He kind of made me feel at home with his twang.”’ Fleischmann, in contrast, revealed an impressive knowledge of neutron detection, although how deep his understanding went was unclear. “He said all the right words,” recalled Palmer. Fleischmann seemed to be hoping that their absence of neutrons could be explained by some loophole in the esoterica of nuclear physics. He thought maybe Jones and his colleagues could, wittingly or unwittingly, help point it out. Fleischmann knew that when two deuterium nuclei fuse the result, in conventional nuclear physics, is either tritium and a proton or helium 3 and a neutron. The probability of either was about

BAD

SCIENCE

# 63

fifty-fifty. Fleischmann repeatedly asked Jones and his colleagues whether this branching ratio, as it is known, could be altered dramatically by the presence of the palladium lattice, which might explain why they had detected some tritium, albeit not nearly enough, and very few or no neutrons. Jones assured him that the branching ratio appeared to be immutable. The fifty-fifty split that nuclear theory predicted was essentially what had been measured in every fusion experiment, including muon-catalyzed fusion. Pons and Fleischmann were disappointed in the answer. Neither seemed willing to doubt their conclusion that they had induced fusion,

at least not in front of the competition. Perhaps their particular brand of nuclear fusion worked in a manner unknown to physicists. Certainly, they had fusion; they simply lacked the conventional evidence—neutron radiation. ““They weren’t dead, for example,” observed Czirr, who knew well enough about the lethal nature of exposure to excessive neutron radiation. ‘‘Although I don’t think they appreciated that point at that tines! The two chemists had brought along one of their own cells, and the BYU physicists were chagrined at how much larger and more professional than theirs it was. After seeing it, Czirr took to referring to the BYU cells as ‘‘the little, dinky baby-food jars.”’ Fleischmann remarked that this cell was just a prototype. It had never produced neutrons, he said, which seems as good a way as any to have forestalled suggestions that they test it then and there. Jones then revealed his neutron spectrum and the infinitesimal bump, which he considered his fusion signal. Pons and Fleischmann did not say what they thought of this evidence. “Oh yes, we showed it to them,” Jones said of his spectrum. ‘““We showed them a bump in the neutron spectrum. It’s a very tiny effect, and I don’t think that sunk in with them.” In fact, Fleischmann did realize it, although he may have construed that the effect was infinitesimal because Jones didn’t know how to construct a fusion cell properly. One week after the BYU outing, Fleischmann called Dave Williams at Harwell and told him that they had spoken to Jones, “who [had] showed them this absolutely grotty neutron spectrum,” as Williams put it. After the tour of the laboratory, it was off to lunch at the Skyroom, BYU’s rooftop faculty club. Now each side’s misinterpretation of the two experiments began to play like farce. Fleischmann, for instance, warned the BYU physicists about the possibility of meltdown. Said Decker later the remark “went right over us.”

64 9 GARY

TAUBES

The BYU physicists then asked Pons and Fleischmann how they measured the excess heat. The physicists were thinking that the heat commensurate with their infinitesimal neutron signal would be equally infinitesimal, and certainly unmeasurable. Fleischmann replied that they

used very sensitive thermometers. And according to Decker the BYU physicists thought, “Boy, they must be really sensitive.” Eventually Jones told Pons and Fleischmann that he was ready to publish his data. He recalled, “‘I had this debate. The scientist in me says,

Look, you’ve been working on it, publish. They don’t have any neutrons. I thought, Well . . .” Jones considered this his deferral to the

Golden Rule. “If they were about ready to publish,” he said, “‘I would want them to tell me. So I told them that, after two years, we’re getting ready to publish.” Fleischmann argued that Jones shouldn’t go public. He was worried that ifJones published the discovery would be lost. Thousands of scientists would flood into the field. Once again the BYU physicists could make no sense of the logic. They knew what kind of neutron detectors the rest of the world had. They thought Czirr’s were as good as any. ““We’re not worried about the rest of the world,’ Decker said. ‘““We

think we are as good scientists as anyone else out there. So we’re not going to worry about it.”’ Fleischmann argued that cold fusion had weapons potential that could be dangerously destabilizing. What if the Soviets developed it first? He did, however, seem seriously concerned with the weapons question, as Ron Bullough had pointed out. (Two months later, however, Fleischmann testified before Congress that he saw no military applications for cold fusion. Perhaps he didn’t want to jeopardize the Utah attempt to procure federal funding for cold fusion research.) In any case, the BYU physicists failed to see the weapons potential of a device that emitted two neutrons per hour. Finally, Fleischmann appealed directly to Jones’s professed sense of fairness. He said he and Pons had worked on cold fusion for years as well, but they needed eighteen more months. If Jones went public before then, he and BYU would receive all the credit. Jones argued that he had been invited to give a talk at the American Physical Society meeting, and he wanted to submit a paper beforehand, and Gajewski at the Department of Energy was telling him that if he wanted DOE funding on cold fusion, publishing a paper would help tremendously. He had no choice but to publish. As an alternative, Jones suggested that Pons and Fleischmann return to

BAD

SCIENCE

@ 65

BYU Monday morning with a working cell. They would test it under Czirr’s neutron detector, and, if neutrons appeared, they would write up the result jointly. Fleischmann reluctantly agreed. Dan Decker later remarked that the BYU physicists simply had no idea what Pons and Fleischmann had observed experimentally or what the two chemists were thinking. As Decker saw it, the BYU results and the Utah results were so dramatically different that the BYU physicists could have published their work and been no threat whatsoever to Pons and Fleischmann. Very few physicists would have even bothered to read their paper. A handful might have tried to reproduce their experiment. No one would have conceived of the BYU-variety piezonuclear fusion as salvation. By Decker’s logic, Pons and Fleischmann could have then published two years later and received all the credit and acclaim they deserved. This line of reasoning illustrated just how little Decker understood, even four months

after the announcement,

about

the ambitions

of Pons

and

Fleischmann and the University of Utah. Paul Palmer, by contrast, had an odd premonition during the lunch. Afterward he wrote in his log: 23rd of February 1989—Visit by Stanley Pons and Martin Fleischmann. U of U. This was a fun day. Pons very quiet. Fleischmann an old time con artist—maybe. At least he is so good that neither Bart, Steve nor I could tell whether or not we were conned. But we knew we’d been conned,

but we didn’t know how.

To Bart Czirr, Monday morning would be the ultimate test. They had known for several months that Pons and Fleischmann had tremendous electrochemical knowledge and no neutron detector. “If I could say there were no neutrons coming out,”’ Czirr summarized, “that would be a very, very strong statement. If it melted my detector, that was also a very strong statement. . . . Either we were going to find a lot of neutrons from their big monster cell or we weren’t going to find any neutrons. ... Their whole program was going to be in big trouble if we didn’t find a whole bunch of neutrons from their big cell.” The BYU physicists all had an initial prejudice. They felt that when they put Pons and Fleischmann’s professional cell in front of Czirr’s neutron detector it would remain as unresponsive as it had when they put their dinky, little baby-food jars in front of it. Nonetheless, they discussed what strategy they would pursue if the Utah cell did emit copious

66 8 GARY

TAUBES

neutrons; their own data would look ‘‘piddling”’ in comparison. They decided that even if the neutrons emitted by Pons and Fleischmann’s cell fried Czirr’s detector, it would be of some historical interest to publish their own data. All of this became irrelevant on Monday, because Pons and Fleischmann never arrived. They were expected at 8:00 in the morning. At ten Pons telephoned Czirr, who had asked Pons to convince one of the Utah physicists to help him with his neutron detector. “We expected you guys down here at eight o’clock,”’ Czirr said. “No, no,” said Pons. “We

aren’t going to be down. We

have a

graduate student who was supposed to set the cell up so we could come down, and his father died over the weekend so he had to go to the funeral. So he won’t be able to set the cell up.”” Pons said he’d call at the end of the week. And that was it. When Czirr relayed the message that the two Utah chemists had stood them up, Jones and his collaborators discussed what to do next. As Palmer remembered it, Jan Rafelski, who had flown up from Arizona for

the occasion, now argued that they had to publish quickly. ““We were floundering around,” said Palmer, ‘‘saying, “Well, let’s see what can we

do to make this work better.’ And Rafelski said, not quite this unkindly, “You dumb guys don’t have a clue as to what you’re doing. You don’t have any theories. There’s no direction. You don’t know what metals to use, what to do next. Listen, if you take all of the foreground and all of the background and add them up, and subtract one from the other, and

the foreground signal is positive, that’s enough to publish on. Let everybody else figure out what you did. Why you couldn’t make it work twice the same way.’ And so he said that, and we said, ‘Yeah, that’s right. Let’s

publish it.” And we decided to publish it.”’ Their goal, as Jones now told them, was to publish as quickly as possible. He said their funding and future research depended on it. The following day, Jones informed the DOE Office of Patent Counsel that he intended to submit patent applications on ‘“‘inventions having to do with piezonuclear fusion.”’

During the following week, Jones and Pons spoke several times. They still discussed the possibility that the two chemists would come down with a cell; then they would write a joint paper that they could publish before the American Physical Society meeting. On February 27 Jones noted in his logbook that he and Pons had discussed obtaining heavy water for the experiments. (Pons: I can get clean D,O.’’) And on March 2:

BAD

SCIENCE@ 67

Pons:

—when coming —yointly publish data acq’d at BYU —Jjoint paper & authorship -Gaj doe plan to get paper in before May mtg (APS)

But Pons and Fleischmann had little Pons’s telephone calls to Jones appear he and Fleischmann determined how The more revealing phone call was

or no intention of collaborating. to have been astalling tactic while to handle the situation. one Pons made to Doug Bennion,

the electrochemist Jones had recruited earlier in February. Bennion, who

had been at BYU since 1980 and had known Pons since the latter had come to Utah, recollected that the call was one of the strangest he’d ever received. Pons wanted to know when Bennion had begun working with Jones. Although Pons didn’t say this, he believed that Jones had recruited Bennion back in September, after reading Pons’s proposal. In fact, Jones at that time had discussed cold fusion with another chemist, James Thorner, who had introduced him to Bennion after Gajewski told him to get an electrochemist. Pons jumped from one question to the next. Bennion later said he felt as if he were being interrogated. Was he working for Jones? Was he building cells? Did he really expect to do electrochemistry for Jones? Yes. Yes. Yes. When Bennion finally got around to asking Pons what this was all about, Pons responded with obtuse warnings: “Be careful! You are going to get hurt; you better not do this.” Bennion felt Pons was trying to frighten him off the project, but the more Pons talked, the more intrigued Bennion became. If this was a way to produce massive amounts of energy, well, then, that was exactly what he wanted to do. Bennion’s

one legitimate concern was that the cells would be dangerously radioactive. When Pons warned him that they could explode, Bennion asked him several times what he meant. ‘Do you mean neutrons are coming off? We are going to get hurt by radiation?”’ Bennion asked. No, Pons said.

So Bennion asked how Pons could possibly be getting energy without neutrons. Pons told him the meltdown story. Bennion later remarked that he worked with batteries and chemicals for a living. Explosions were an unfortunate aspect of his life. So what? Once again, he asked how they

68 8 GARY

TAUBES

could be seeing a nuclear reaction without radiation. Pons only warned Bennion once again to be careful; then he hung up.

The outing at BYU strengthened Pons and Fleischmann’s.convictions on two points. It confirmed their suspicion that Jones had pirated their theory for inducing fusion. They reached this conclusion because Jones, a physicist, was indeed doing electrochemistry, not to mention doing it with palladium electrodes, just as they were. (It probably helped that Jones kept referring to his two or three years of concerted effort, while Pons and Fleischmann saw an amateurish experimental setup that couldn’t have represented more than a few months of work.) Why would a physicist think of doing anything electrochemical? From a chemist’s viewpoint, it looked as though Jones had thrown together an experiment that made no chemical sense and was trying to scoop them. More important, the visit to BYU

seems to have further persuaded

Pons and Fleischmann that they had generated room temperature fusion. Their reasoning seemed inarguable: their cells had to be radiating neutrons because Jones’s baby-food jars were radiating neutrons, and Jones was employing a technique that they had pioneered. If the electrolysis technique did not lead to nuclear fusion, why else would Jones be working so diligently to steal it? And why else would he be insisting that he had to publish it? What could be more self-evident? University of Utah president Chase Peterson came to embrace this logic and boiled it down to its essence: “She must be a good-looking girl,”’ he observed, “‘if somebody else wants to date her.”

CHAPTER FEBRUARY

26TO

5 MARCH

15,

1989

hen Pons and Fleischmann returned from BYU, they reported to Jim Brophy, Utah’s vice president for research, that their cold fusion work had been leaked. Brophy explained to Chase Peterson that the leak had traveled from the Department of Energy to Brigham Young University. Peterson was understandably unhappy, but he was prepared to take control. Cold fusion was no longer just a scientific issue; it had become to Peterson an administrative one as well. Peterson had been in a bind since he became president in 1983—he had the highest aspirations for his university but limited resources. Given those circumstances, he may have been ripe for cold fusion. He’d had his eye out for salvation, and once he believed he had seen it, he wasn’t

about to question it, or entrust it to anybody else. Peterson’s ambitions for his school—and his state—grew out of the fact that he was the product of two cultures. On the one hand, as he noted at his inaugural convocation, he had deep Utah roots. He was born

there in 1929, when his father was president of Utah State Agricultural 69

70 BE GARY

TAUBES

College, now Utah State University. Being raised on a college campus informed Peterson’s life with a devotion to the educational process. At age fifteen he won a scholarship to Middlesex School in Concord,

Massachusetts. This may have been a defining experience, for he spent twenty-nine of the next thirty-four years living and working with those he would later refer to as the Eastern Elite. One of his colleagues suggested that “‘a scholarship student from the sticks of Utah attending such a hoity-toity institution would be affected for life.”’ After Middlesex,

Peterson went to Harvard

and Harvard Medical

School. He did his internship at Yale-New Haven Medical Center and returned to Salt Lake City for all of five years in 1962 to teach and practice in a clinic. But the lure of the East won out again. Back he went to Harvard for eleven more years, first as dean of admissions, then as vice

president of alumni affairs. It wasn’t until 1978, at age forty-nine, that Peterson returned to Utah to become vice president for health sciences

at the U.

After March 23, Peterson fell back on his belief in the Eastern Elite’s

superiority complex to rationalize the shellacking he received from the more prestigious scientific institutions, which he insisted were just “turf protecting.” No one had data disproving Pons and Fleischmann, he would insist; it was simply a battle between “hinterland science and science on the coasts.” This delusion reveals what Peterson had in mind for cold fusion all along. He wanted to transform Utah and the university into an intellectual and economic force to be taken seriously. Indeed, this ambition to be taken seriously seemed to be endemic to the state of Utah and fed an institutionalized insecurity complex. Rod Decker, alocal political pundit, put it this way: “If you aska lot of people in Utah what they fear, it’s that everyone is going to laugh at us. We want money and we want respect. We would very much like to be respected, not as Nevada is for gambling, or Wyoming, for coal, but for brains and

for talent.”” Thus, Peterson, two months after the announcement of cold

fusion, was quoted in The Salt Lake Tribune saying that his university was in a fight against what he called the Utah effect, “which causes people to believe ‘This may be important, but it may not be believable’ because it came from Utah.” The state’s commitment to intellect reveals itself in the fact that, per capita, Utah had the most college graduates, advanced degrees, and

Ph.D.’s in the nation. But the locals still realized that, if their state were

known as anything, it was as the home of the Church of Jesus Christ of

Latter-day Saints, the Mormons. After cold fusion broke, for instance,

BAD

SCIENCE

@ 71

the London Sunday Telegraph referred to Utah as “‘a scenic and empty state whose best-known contribution to Western culture until now has been the Mormon Tabernacle Choir.” For 150 years the local population had been living with the irksome fact that their church was viewed by the outside world as nothing more than a particularly wealthy and successful cult—and one with more than its fair share of exotic rituals, including, of course, polygamy. Lest the people of Utah be allowed to forget, Brigham Young, “‘the American Moses,” had had twenty-four wives. Although the church had officially disavowed polygamy at the turn of the century, the practice still lived on, to the chagrin of many, in backwaters of the state. The last front-page news to come out of Utah before cold fusion was in January 1988, when

the Singers, a family of radical Mormon polygamists, held off police in a bloody and lengthy shoot-out. In the words of historian Charles S. Peterson (no relationship to Chase Peterson), “Most Americans judged Mormons first unique, then foreign, then alien, then the enemy of American life. Mormonism seemed a counterculture that rejected America’s progress, harked back to intolerance, mixed church and state, and

threatened property.” A less specific variant of this judgment has been proposed by Paul Fussell, an Eastern Elite literary and cultural critic. In describing the class structure of America society, he wrote off Utah thus: “Indeed, it seems

a general principle that no high-class person can live in any place associated with religious prophecy or miracle, like Mecca, Bethlehem, Fatima, Lourdes, or Salt Lake City.”

The recent history of Utah hadn’t managed to ameliorate the statewide image problem, or the insecurity of the state’s residents. Standing in the lineup alongside the Singers was Mark Hofmann, of the Salamander forgery trial. Hofmann’s penchant for letter-bombing his associates exposed him as a forger who had sold fake Mormon documents to highly placed members of the church. Then came the story of Gary Gilmore, the local murderer who was executed in 1977 and immortalized in Norman Mailer’s book The Executioner’s Song. Salt Lake City had also earned some lasting fame as the “‘scam capital of the country,” because it generated much more than its fair share of securities fraud. Utah seemed to be cursed as the home of some of the newspapers’ most sensational crime stories. One shining exception to this public relations nightmare was Barney Clark and his artificial heart, and that had been Chase Peterson’s show.

He had done a tremendous job, but he still couldn’t win outside Utah. He was swinging for the fences, said his critics, trying to use Barney Clark

72 @ GARY

TAUBES

to turn Utah into the “‘bionic state,”’ the artificial organ equivalent of Silicon Valley. Rarely, however, did they think of it that way in Utah. Peterson’s handling of the Barney Clark episode increased his standing as a candidate for the presidency of the university, although he would have been a strong candidate in any case. (When a New York Times reporter asked Peterson what role he might tackle next, ‘‘he shrugged his shoulders and said perhaps a tour of duty as a Mormon missionary in Bolivia. Perhaps politics. In 1981, only three years after he returned to Utah, Dr. Peterson’s name was mentioned in Democratic circles as a potentially attractive nominee for the United States Senate.” He had proven himself to be a good administrator and a promoter of his university. He proudly referred to himself as a “‘booster.’’) He was also a Mormon, an implicit prerequisite for the job. And he was neither overly righteous about his religion nor dangerously iconoclastic. When cold fusion came along, Peterson had been president of the university for five years, and it had not been a smooth ride. He had shown atalent for fund-raising and lobbying. He had cemented solid relationships with the governor and other members of the political leadership. But whereas his predecessor, David Gardner, who had left to take

control of the University of California system, had run the U when the worst problem in the state seemed to be how to spend all the surplus revenue, Peterson was not so lucky. The fortunes of the state of Utah tend to go up and down with the price of oil. The Utah economy had spiraled up with Gardner and had spiraled downward with Peterson. When energy prices dropped in the 1980s, Utah was hit almost as hard as Texas. The market for timber and

metal mining, two usually bankable industries for the state, also deteriorated in the eighties with the rise of the dollar. Two of the largest employers in Utah, Kennecott Copper and Geneva Steel, shut down entirely for two years in the midst of Peterson’s reign.” Since 1981, according to The Salt Lake Tribune, Peterson had been forced to cut his university budget on eight occasions. In 1986, the worst year of all, Peterson had to cut $9 million and 200 jobs. Many professors had not had a raise in four years. They were earning, on average,

20 percent less at Utah than they would have gotten at comparable

Institutions.? Some of the more prestigious faculty members had moved on, including William DeVries, the surgeon who implanted Barney Clark’s artificial heart. (Even after the cold fusion announcement, Peter-

son appeared overly anxious that he would lose his two cold fasion

chemists as well. “We don’t own Dr. Pons and Dr. Fleischmann,”’ he

BAD

SCIENCE

& 73

said. “‘They are independent. They could go to Switzerland. They have been invited to alot of places.’’) In the fall of 1988, the year before cold fusion, Peterson barely managed to head off a statewide tax reduction initiative that would have cut his funding by another 11 percent. Locals recalled that, at one point during the year, he took to carrying out his own garbage to set a frugal example for his faculty and employees. So Peterson was more than primed when Jim Brophy told him about cold fusion in the fall of 1988. Ed Yeates, a local television reporter, was

doing a story on the university’s latest financial plight when Peterson took him aside and told him that the U had something big in the works, but he couldn’t say what it was. As Yeates recalled it, “He said, ‘I don’t

know when it will happen, but it will be soon. And it involves chemistry.’ ”’ Peterson also mentioned that it could lead to Nobel Prizes. It was rumored that Peterson told the governor, Norm Bangerter, essentially the same thing in December or January. Also about this time, perhaps late December, Brophy announced to the president’s cabinet that Pons and Fleischmann had come upon a new way of creating fusion. Jim Brophy’s role in cold fusion was crucial. He appeared to be incapable of doubt. “If you come down to the question of doI believe there is science there,” Brophy would say, “‘my answer has always been yes.” As the story evolved after the announcement, Brophy seemed willing to report whatever he felt was necessary to sell cold fusion or, in his words, ‘‘to put the best possible honest face’? on any cold fusion— related development.* Before cold fusion came along, the faculty at Utah seemed to have nothing but respect and fondness for Brophy. He was considered invariably straightforward and honest. He was known asa facilitator. In a job that was usually the provenance of administrators obsessed with proper protocol and bureaucratic consideration, Brophy could get things done. He had come to the U from the Illinois Institute of Technology during Gardner’s tenure, and Peterson had kept him on. Brophy was one of the few non-Mormons in the administration. Joe Taylor said one reason he liked Brophy was that at get-togethers at Peterson’s house, “neither one of us is afraid to have a drink.” Indeed, Brophy, who grew up in and around Chicago, was a childhood friend of Hugh Hefner. Hefner had been the best man at Brophy’s wedding, and when cold fusion broke, Hefner wrote to Brophy, as one of Pons’s chemists remem-

bered it, ‘‘congratulations on this fusion, etc. As you are aware we've been conducting our own fusion research here in the mansion for some time.” And when Hefner got married, Brophy apparently sent him a

74 © GARY

TAUBES

nineteen-dollar toaster and two cold fusion T-shirts, which had become

hot commodities at the Utah bookstore. (One Utah administrator later suggested that cold fusion was Brophy’s gift to the scientific community, the equivalent of a nineteen-dollar toaster.) Brophy’s faith in cold fusion helped convince Peterson of its potential, and Brophy gave weekly and uncritical progress reports on the research at cabinet meetings. The cabinet members were sworn to secrecy, and

in the minutes the word fusion was never used. It was referred to as the “F-word.”

Peterson’s first administrative decision in the escalating cold fusion crisis was to minimize the damage done by the leak of information to BYU. He called Norm Brown, director of the U’s Office of Technology Transfer, on March 2 and requested that they file a cold fusion patent as soon as possible.* Brown in turn called Peter Dehlinger, a patent attorney in Palo Alto, California. Dehlinger had a doctorate in biophysics from Stanford and had worked in the field for five years before picking up his law degree. The two men had worked well together in the past, and Brown wanted Dehlinger to handle the patents because he also thought the U would need the expertise of a physicist. Dehlinger readily agreed to take on the assignment, then spent the weekend in the library “trying to figure out fusion.” On Friday, March 3, Peterson phoned Jeffrey Holland, the president of BYU, to arrange a summit conference between the two universities. When he couldn’t get through to Holland immediately, he went to Jae Ballif, who was provost and, coincidentally, the first cousin of Peterson’s wife. The two families were very close. As Ballif remembered the ensuing conversation, ‘‘He talked to me about the need for us to meet together and asked if I was aware of this research going on. I told him I knewalittle bit about it but not much. Chase said that there was some question about provenance of ideas, and he informed me for the first time that they had people there doing something on the same subject. And he said it’s of gigantic proportions in terms of its economic and scientific importance, political importance, and we needed to give it first priority and meet together.”’ Ballif agreed. He reached Holland and arranged the meeting for first thing Monday morning. By all accounts, Peterson went into this meeting planning to come to a diplomatic and equitable agreement. The two schools had successfully collaborated in the past, and he had no reason to doubt that they could do so again. To Peterson, this spirit of cooperation was the Utah way;

BAD

SCIENCE

@ 75

indeed, he believed it was unique to the state. This was the flip side of the Mormon experience. The locals liked to say that they knew how to circle the wagons. This cohesiveness arose not so much out of the Mormon religion but in response to the years of persecution. The Mormons had fought the army of the federal government over the right to exist in 1857 and fought the government again forty years later over the polygamy issue before they could obtain statehood. More than anything, from the nineteenth century onward, the Mormons had fought for respect. After setting up the BYU meeting, Peterson went to recruit Joe Taylor. As Taylor recalled it, Peterson was obviously agitated and argued that a public announcement by Steve Jones would cut the ground out from under their own effort. He said Jones was offering a collaboration, but Pons and Fleischmann would have none of it. If all parties could sit down around a table, which was what Peterson had planned, he was sure

they could work out a deal. If they couldn’t settle on a collaboration, they would at least agree to postpone any public announcements, then make a joint announcement when both sides were ready. This strategy struck Taylor as a mistake, even at the time. “I didn’t know why I thought it was a bad idea,” he said, “except it didn’t seem like the kind of thing that would lead to anything good.” Peterson wanted Taylor to accompany him to BYU as a witness to any agreements. Brophy, who would have been the natural choice, was going to be out of town on Monday, so Taylor agreed. After Peterson left, Taylor discussed the situation with Brophy, who confirmed Peterson’s account. Taylor said he thought Peterson’s strategy was a bad one, and Brophy agreed. ‘Do you think we should talk him out of it?” Taylor asked. “No chance,” Brophy said. “It’s all set up. He won’t back out now.” At BYU, Ballif remained baffled over Peterson’s statement “‘that there

was some question about provenance of ideas.” Ballif went to Grant

Mason, dean of the College of Physical and Mathematical Sciences, who

briefed him on what he knew of the controversy between Pons and Fleischmann and Jones. The two then called in Jones, who gave his version of why Peterson might have been so anxious to hold the summit meeting. Jones assumed that Peterson was concerned about whether or not Jones had begun his cold fusion work independently. Ballif found Jones “‘so believable and so open” that he decided the best way to get Peterson to understand the absurdity of any possible piracy accusations was to hear Jones himself talk about what he’d been doing.

76 0 GARY

TAUBES

Already Ballif and Jones were misinterpreting the reason for the summit conference. (Indeed, Peterson hadn’t told Ballif exactly what he had in mind, only that they had to meet and discuss it immediately.) All the same, Ballif suggested that Jones gather the documentation of his history of cold fusion work, so he could present it Monday morning and clear the matter up. Paul Palmer and Bart Czirr were pessimistic about the meeting. Palmer’s notebook entry for March 6 reads: Bart said Pons and Fleischmann are going crying to the president for him to come—to bear down on us and stop us before we get beyond our discovery of the heat engine that drives plate tectonics—and may power industry. They don’t want us to get the Nobel prize, when it is in their grasp.

Czirr also predicted that it would be bad business to tell Pons and Fleischmann when BYU had begun its cold fusion research, because the Utah scientists would concoct a scenario in which they had begun one or two years earlier than Jones had, which appeared afterward to be what happened. During the forty-five-mile drive from Salt Lake City south to Provo on the morning of March 6, Pons and Fleischmann talked about their experiment. Taylor later remarked that he learned more in that drive than he had during the two months of Brophy’s briefings to the president’s cabinet. Pons and Fleischmann talked about the meltdown, and

Taylor and Peterson were suitably impressed. Fleischmann said, as he had told Jones two weeks earlier, that they would have liked another eighteen months to pursue their research because the abundance of tritium and neutrons they had measured was not sufficient to explain the magnitude of the heat they had seen. Neither Taylor nor Peterson thought to ask how great this discrepancy was. Taylor later said that he wished he had. The summit meeting began shortly after 9:00 a.m. in Room 301 of the BYU Administration Building. Picture windows framed a view of the mountains. On one side ofa large table sat the BYU contingent: Jeffrey Holland, who was soon to leave his presidency to become one of the Seventy, which was the third level of leadership of the Mormon church; then Ballif; Lamond Tullis, who was the associate academic vice president and a Harvard-educated political scientist; and Steve Jones. Across

BAD

SCIENCE

@ 77

the table, facing the windows and the mountains beyond, were Pons, Fleischmann, and Taylor. Peterson conducted the meeting and stood through most of its three hours, often at the blackboard at the head of the table. The purpose of the meeting, he announced, centered on his hopes that the two universities could come to a managerial, social, and political, if not scientific,

agreement on cold fusion. While he spoke, he chalked little boxes and arrows on the board to illustrate his points or glanced down at his scribbled notes to keep his place. Peterson proceeded to lay out the dimensions of the cold fusion property: the potential of a limitless, clean source of energy in a world beset by environmental problems and dependent on Middle Eastern oil. He described nuclear fusion and how it differs from nuclear fission; he

even sketched the two mechanisms on the blackboard. Tullis couldn’t help but wonder why Peterson had gathered all this high-level administrative power only to make them sit through an elementary physics lecture. Even if they weren’t scientists, they were all literate. Peterson moved on to the huge sums of money involved in cold fusion. How could they assure that they developed this technology in Utah and didn’t lose it to the technological wizards in Boston or Palo Alto or, worse, Japan? He discussed the impact the potential wealth of OPEC would have on the local and the national economy. He even broached the weapons issues. After all, cold nuclear fusion appeared to produce radioactive tritium, which is a necessary ingredient in the production of nuclear weapons. Peterson said he believed that a satisfactory agreement between the Y and the U could be reached. He sketched out his views on how the scientific credit could be shared, hoping that his two chemists and the

BYU physicist could publish simultaneous articles announcing the discovery. He hoped that they could work out the issue of patents and future research support. This was critical because Nobel Prizes were at stake, not to mention billions and billions of dollars.

To the BYU contingency, Peterson’s discourse seemed to roll on interminably. Tullis’s thoughts wandered over the question “How is it that Nobel Prizes are won?” Are they won by the development of a political agenda or on the basis of a substantive contribution to the science of our times? He was not so naive that he didn’t know it was some of both, but...

Eventually Peterson brought his sprawling presentation to an end.

78 © GARY

TAUBES

Ballif then said, “I think we should hear from Steve.” Jones used his own hour to cover his work in cold fusion. First, however, Jones wanted to respond to Peterson’s talk of the

billions of dollars to be made from cold fusion. As Jones saw it, cold fusion was only a scientific curiosity, certainly nothing that would save the world. He reached into his small blue bag and retrieved a tiny flashlight. Flipping the light on, he said in his soft voice, ‘Look, I don’t mean to be rude, but we’ve been looking at this process for years now, and it’s just not an energy producer. If you could ever get enough energy to light a flashlight, I’d be extremely surprised.”’ Fleischmann then countered that they had generated enormous amounts of heat from their fusion cells and had had an explosion. Fleischmann may have been warning Jones that they were playing with dangerous stuff, but Jones didn’t buy it. This further convinced the Utah contingent that Steve Jones, physicist though he might be, was ignorant of both the science and the potential of cold fusion. Jones’s insistence that cold fusion was not worth the trouble to patent made Joe Taylor momentarily optimistic. Obviously they were talking about two kinds of phenomena. Jones must have been seeing something quite different from what Pons and Fleischmann were dealing with. Maybe he wasn’t performing the experiment correctly, certainly not the way Pons and Fleischmann were. None of the assembled administrators seemed able to grasp the fact that the effect Jones believed he had observed and the effect Pons and Fleischmann believed they had observed differed from each other in magnitude byafactor of one trillion. Jones would later take to describing the difference as though he had observed an effect comparable in power to a single dollar bill, and Pons and Fleischmann were talking about an effect comparable to the entire U.S. national debt—$1 trillion. Then Jones began patiently to document his personal history with cold fusion. This was the same account that the scientific community would hear frequently in the next few months. Jones made a special point of displaying his lab book entry from April 7, 1986, on which, he said, he had discussed the very idea that Pons and Fleischmann had accused him of pirating. This page was notarized. It was signed by BYU’s patent attorney. Notarized? Even Ballif, the one who had asked Jones to document his history, was

stunned to see that Jones had thought to notarize his notebook. Nobody did that. Still, it seemed to make absolutely no impression on the Utah chemists.

BAD

SCIENCE

#79

Finally Jones came to the end of his history lesson, and Peterson got to his real agenda. He asked that Steve Jones refrain from publishing his research. He said Pons and Fleischmann needed another eighteen months to do their seminal experiment, and at that point they could all publish simultaneously. No one apparently thought to ask Pons and Fleischmann how they could be so confident that they had discovered anything at all if they needed eighteen more months to prove it. Ballif, for one, assumed that,

like all scientists, they simply would have liked time to amass additional evidence.® Jones explained that he had been invited to give a talk on cold fusion at the American Physical Society’s spring meeting on May 3. Peterson asked him to cancel the talk. IfJones spoke before Pons and Fleischmann were ready, and before all the possible patent issues were finalized, his talk would constitute what patent attorneys call a public disclosure. The last thing Peterson wanted to see was a premature public disclosure of cold fusion from BYU. Such an act would assure that anyone in the world could steal the Utah discovery and profit from it. Utah could end up without a dime. But Jones refused to yield on this point.” He said that he was clearly ready to publish, and that the Department of Energy, which had paid for the work, was virtually forcing him to talk about it. In fact, he showed the abstract describing his talk that he had sent off to the APS offices a month earlier. Jones thought he was being open and honest, showing and telling the Utah contingent everything. The abstract said he had “accumulated considerable evidence for a new form of cold nuclear fusion.” Now Peterson seemed truly worried, which even Jones noticed. “Are you serious that you really need to report this right now?” Peterson asked. ‘““What if more time would allow this thing to mature more importantly, for the universities, for the state?” Jones said he had no choice in the matter. Peterson couldn’t talk him out of it. Jones’s intransigence made Pons and Fleischmann furious. Cancel an invited lecture? No, no. I couldn’t

possibly. The two chemists were thinking, What’s so important about an invited lecture? We get requests to speak all the time. Jones then added that he was scheduled to give a campus colloquium on cold fusion in just two days. As Peterson recalled it, he responded, “‘Gosh, Steve, you understand

this may have some importance, and if there’s a student reporter who happens to be taking physics that day and wanders in, that can be put in

80 BF GARY

TAUBES

the student newspaper and then it can go to the national pickup and now you've lost your patents.” Peterson believed that he was talking very openly. If he was trying to protect patents for his university and cheat the BYU physicists out of what was justly theirs, he would have been much more cunning and consciously avoided the subject. But every time the BYU group insisted that they were not interested in patents, he insisted that they ought to be. Peterson felt that the best possible outcome was for the two universities to pursue cold fusion as a joint venture: “But they were not showing interest,” he noted later. “‘[Their position was that] this is all for the good of mankind and so forth. I wanted it to be for the good of the state of Utah.” Jones agreed to cancel the colloquium, but he would not back out of the invited talk in May. They turned to the question of a collaboration. All the outward civility, however, could not mask the fact that Pons and Fleischmann

were in no mood to have Jones in their laboratory and vice versa. The conversation worked back around to Peterson’s original idea of simultaneous publications. Pons and Fleischmann needed eighteen months, but now May 3, the date of Jones’s invited talk, was the outside deadline. If the Utah chemists wanted to get the word out and share the credit, they had to do it by then. In fact, Jones’s abstract would appear in print the first week of April. Peterson finally suggested that they publish back-to-back articles in a journal, but Fleischmann said they were not sure that they could produce a definitive paper in time. Pons said nothing. He barely uttered a dozen words through the entire meeting. Jones suggested they publish in a physics journal, in particular the most prominent, Physical Review Letters. Fleischmann thought this was too abstract. He said his preference would be to publish in Nature, a prestigious British journal. Should Nature not work out, he knew of an electrochemical journal that would surely accommodate them. But Jones argued that the BYU work, at least, would not be appropriate for such a narrow chemical journal, and he preferred to publish in what he later called “some reputable, refereed journal.’’ He would not concede the point. After all, he was the one who was ready to publish. They let it go at that. Jones would talk to the editors of Physical Review Letters, and maybe Nature as a substitute should Physical Review Letters be unable to publish fast enough. Peterson then insisted that it would be unconscionable for Jones to

publish or discuss the research before this back-to-back submittal. With

BAD

SCIENCE

@ 81

Jones having agreed to cancel his campus colloquium as well as a scheduled talk by one of his students, the BYU delegation thought that they had been both generous and accommodating. Later Peterson said that his people hadn’t agreed to publish simultaneous papers so much as they had acquiesced to the idea. Joe Taylor, who had remained strictly an observer, couldn’t help but think that they had spent too much time on pleasantries and never allowed their real feelings to show. Here were the presidents of two neighboring universities bending over backward to be polite to each other. It was Peterson’s show, so Taylor didn’t intrude, but he later

wished that he had taken Peterson aside or requested a ten-minute break. If they had just gotten the chance to discuss the situation alone, he thought, they would have realized that it wouldn’t work. By contrast, Lamond Tullis, whose field of study was the politics of international narcotics trafficking and who did not appear naive, thought they had been in perfect accord. When the meeting broke, they shook hands all around. “‘That’s part of the local custom here,” Tullis explained. “‘[And] you don’t shake hands vigorously with people with whom you still have substantial disagreements without noting that those disagreements exist.” So after three hours of near total misunderstanding, the men strolled across the BYU campus for lunch at the faculty club. (Jones excused himself, saying he hada class to teach.) As they walked, Pons mentioned the meltdown he’d accidentally induced in his laboratory. He told Tullis that a palladium cube had melted through four inches of concrete. It was one of the few things Tullis could remember Pons saying. Over lunch the group listened to Fleischmann tell stories of his childhood. Both Tullis and Ballif found him tremendously congenial. They wondered about Pons, however, who was still so quiet. Perhaps there was no way that they could relate to Pons. Tullis later realized that Pons must have felt surrounded by a bunch of thieves and crooks. Taylor was uncomfortable as well. He was beginning to feel like a player in some absurd farce. An appalling situation was in the works; they had come to a compromise that represented only Utah’s absolute failure to negotiate. Pons and Fleischmann were obviously seething, yet they were all having a pleasant lunch and telling stories of the good old days. It was surreal. By the time the Utah contingent walked back across campus to the parking lot, Pons and Fleischmann had begun cursing Jones: he was a scoundrel; he was trying to run off with their work. As Pons, Fleischmann, Taylor, and Peterson drove back to Salt Lake

82 8 GARY

TAUBES

City, the chemists insisted that Jones’s hour-long documentation of his research in cold fusion meant absolutely nothing. Peterson later recalled that one of the chemists, he didn’t remember which, said, “Yes, when

I was a boy, I thought of going to the moon. That doesn’t mean I was doing rocket research before Goddard.” All four concurred that the newly minted agreement with BYU was unworkable. They had negotiated themselves into a no-win position. As Pons and Fleischmann saw it, any way this simultaneous submission deal played out, Jones would get more than he deserved, and they would get less. They still believed that Jones had no right to publish his electrolysis results, and now they had traded away any leverage to stop him. For instance, if Jones simply published his geophysical evidence for fusion, which was the only aspect of the business that was clearly his, then their

“more cosmic and important paper was going to be given the same kind of treatment that a minor geochemical paper was given.” They would have to conceive another strategy. ‘‘I guess we have no choice,”’ said Fleischmann repeatedly. “This is getting to be impossible to contain. We are going to have to make an announcement before too long.” Pons, Fleischmann, Taylor, and Peterson discussed the possibility that Pons present the data at an American Chemical Society meeting in Dallas, which would be the second week of April. That would give them a month to reconcile the discrepancies in their data and put together a paper. They discussed calling BYU and telling them the deal was off, but Peterson felt that such an act could prompt Jones and BYU to go public immediately. He felt they couldn’t take that chance. Peterson’s attempt at diplomacy was looking increasingly dismal. As he drove the two chemists home, Peterson was thinking, What in the blazes do we do

now?

The word that the summit conference had been less than a success trickled back slowly to BYU. Initially, Jeffrey Holland told Ballif, who told Jones, that they had done a good job and that “this is going to be a wonderful thing—the two universities working together.” Still, a few of the administrators remained puzzled as to why the meeting had taken place. John Lamb, the director of research administration, raised this question with Lee Phillips, the director of technology

transfer. Lamb wanted to know if they had jeopardized their patent position by having Jones reveal so much about the history of his work. Phillips, likewise, was suspicious about why the Utah contingent would

BAD

SCIENCE

@ 83

want to dig into their lab books. Neither was aware that it was Ballif and Jones who had been anxious to present the data from the lab books, that the Utah contingent hadn’t cared if they heard it or not. Several days after the meeting, Lamb bumped into Jim Brophy at a meeting in Salt Lake City and asked about the consensus at the U regarding cold fusion. “Jim stated point-blank,”’ Lamb said, ‘‘that everyone was convinced that Steve Jones was stealing Pons’s ideas from the proposal, and that they were pretty darned upset about it.’’ Lamb characterized Brophy’s manner as ‘‘quite exercised.”’ When Lamb returned, he wrote a memo to his superiors, noting that, in light of Brophy’s remarks, the administration at the U might be prone to do anything. He suggested that they tread very cautiously, being careful not to reveal more than they already had about their own patent position. He also suggested that all subsequent contacts with the University of Utah be put in writing. The administration, however, chose to take a more benign, wait-and-

see attitude. Holland evidently felt that his university had made certain commitments

and should follow through on them, in Lamb’s words,

“come hell or high water.”’ The official position of BYU was that the meeting had been a complete success, and that their scientists had been the ones who had bent over backward “‘to accommodate the University of Utah delegation.” The official BYU history of cold fusion later enumerated the agreements reached in the meeting as follows: first, that the two sides had agreed to prepare and submit articles simultaneously to the same journal, assuring that ‘‘every effort would be made to get publication prior to the American Physical Society Meeting, even though this would be difficult in the short time available.’’ And, second, that ‘‘no further public an-

nouncements of the results of either team’s research would be made until after the papers were submitted for publication.”’ The history made note of these two points because the University of Utah would violate them both. In accordance with the latter requirement, Jones had agreed to cancel the department colloquium scheduled for March 8, at which he, Palmer, and Czirr had planned to talk. “They stifled the announcement of an . important scientific discovery,” Palmer said.

In the weeks leading up to March 23, if one chooses to believe Steve Jones, he never doubted that he was ready to publish his cold fusion work. After the fact, Jones said he could not recall trying to beat Pons and Fleischmann to publication or trying to force the issue by claiming

84 8 GARY

TAUBES

he was ready to publish when he was not. This is possible, but the data he eventually published contradicts this account, as do his logbook entries. As of the March 6 meeting, Jones’s evidence for the existence of cold fusion was twofold: he had the geophysical evidence, which he and Palmer considered compelling but only because they were not professional geophysicists, and he had the neutron data, which may have been even less compelling. On March 7, the day after the summit meeting, Jones noted that he

had called Jan Rafelski, who had suggested that they run one more set of fusion cells, and that one such run can be significant by itself. He also suggested that they do a quick analysis, publish, and take their chances. Jones wrote, “If4 sigma publish.” Sigma is a measure of the statistical significance of a signal, which is to say, the odds that the signal is real as opposed to some rare fluctuation of the background. It is known as the confidence level. Physicists talk about a three-sigma event or a six-sigma event, for instance. The higher the sigma, the more confidence that the effect is real. According to the statistics books, a one-sigma signal, for instance, has a 31.67 percent chance of being wrong compared with a five-sigma signal with a 0.000058 percent chance. It should come as no surprise, however, that statistics seldom reflect

real life, so they have to be adjusted downward. These probabilities assume that one understands the experiment perfectly, and knows and has accounted for every possible source of error, which is an impossibility. Thus, what is known in the business as the Vernon Hughes law of low-level statistics, named for an atomic physicist of note at Yale. His law states that despite the fact that a three-sigma effect appears to have a 99.73 percent chance of being right, it will be wrong half the time. This is real life. Quite a few physicists would consider publishing a five-sigma event, but most such signals that make it to publication turn out to be delusions. Publishing a four-sigma event is either foolhardy or an act of desperation. One thing was certain: no one who hada four-sigma signal was ready to publish. After March 6, when he began writing his papers, Jones did not run any cells. Between then and March 23, Jones and his friends were writing and analyzing the data they already had, which is to say attempting to prove that the meager signal was real, that the number of neutrons they had detected when the cells were running was undeniably larger than the

BAD

SCIENCE

@ 85

number detected when they were not running. By the time Jones finished his paper on March 23, his neutron signal was considered believable primarily because it was infinitesimal. Meanwhile, Jones and his colleagues investigated which journals might accommodate simultaneous submissions and rush them into publication in the same issue. Dan Decker called a friend at Physical Review Letters and was told that it would be difficult to publish by May 1, because the papers would have to be sent out to referees. Jones phoned Laura Garwin, an editor of Nature with whom he had worked a few years earlier on an article on muon-catalyzed fusion. Garwin told him that she couldn’t promise anything on such short notice, ‘‘and she wasn’t wild about back-to-back submissions, but she said she was willing to work with us.” ““Fever-pitched”’ was how Marvin Hawkins described the first weeks of March in Pons’s laboratory. The fusion cells required monitoring twenty-four hours a day, not for safety purposes but to record continuously the voltages and currents going into the cells and the heat coming out. When Hawkins’s stamina ran out, Fleischmann or Pons would come

in and take their shifts.* Fleischmann had said again and again that they needed eighteen months, at least, to finish the necessary experiments. They discussed the time constraint frequently, asking each other the same questions: ““What are we going to do? If we have to press for publication, what matrix of data can we collect to make a compelling argument for fusion?” Hawkins remembers that they ran the calorimetry experiments until they were convinced that the cells were producing too much heat to be explained by any chemical means. Then they would take the cells apart and rebuild them and run them again until they gave the anomalous heat or burned out. Some ran for as little as a few hours before they died, and others that had been running since Christmas continued to generate more heat than the conventional wisdom dictated was possible. The data, Hawkins said, ‘“were coming off quite reasonably. There were times when I built cells and had seven out of ten work.”’ Still those cells obstinately refused to emit neutrons. Fleischmann had already called Dave Williams at Harwell and explained the problem to him. The neutron detector that Fleischmann had purchased from Harwell, as Williams recalled, ‘‘should’ve gone berserk, and they weren’t

seeing anything on it.” Fleischmann had also asked Williams if Harwell was prepared to put

86 8 GARY

TAUBES

together some fusion cells and count neutrons for them. Williams went to Ron Bullough, the lab’s research director, with the request. “Just do it,” Bullough said. On March 9, Fleischmann faxed Williams instructions to make two of

their fusion cells. One was the cell design that Hawkins said he had conceived in November. In his note, Fleischmann referred to it once as

the “‘awful cell’ and then wrote, ‘Very bad geometry for counting?’’ So Fleischmann apparently believed, or at least hoped, that fusion was occurring and neutrons were being emitted but that somehow the configuration of the cell prohibited their appearance or detection. Fleischmann also included much less detailed instructions for building what he called “‘our No. 2 monster.’’ And here Fleischmann noted, ‘“‘Much better

geometry?”’ Perhaps the monster had seemed less reluctant to emit neutrons. At Harwell, Williams called David Findlay, a Scottish physicist who had been recruited to help with the neutron detection. Findlay said he had this ‘“‘tremendously sensitive neutron detector” up and running, and he was ready for cells. He was not optimistic. “I didn’t believe it at all,”’ he later recalled. “Somehow the idea that you get a beaker of heavy water and a few metal electrodes, connect a battery, and, bingo, neutrons

come flashing out. . . . I mean, people have been trying to produce neutrons for a long time.”’ Findlay’s partner, Martin Sené, said that when he saw the fusion cells his first reaction was “‘hysterical laughter.” Despite the physicists’ skepticism, Williams had the Harwell technicians build fusion cells to match Fleischmann’s drawings and took them over to Findlay’s neutron counter, only to find that they didn’t fit. “There was some grinding of teeth over that,’ Williams said. After alterations the fusion cells were inserted gently into Findlay’s detector and set to charging. It was the beginning of an exasperating wait for neutrons at Harwell. Pons, meanwhile, was procuring physics expertise closer to home. He called the Nuclear Safeguards Division at the Los Alamos National Laboratory and spoke to a division director, George Eccleston. Pons asked what kind of neutron detection capability they had. Was it portable? Specifically, could it be transported up to his laboratory? The answer was yes. But Eccleston also said that he’d need to know about source strengths and setups, and it would take longer than Pons expected. “You can’t go in overnight,” Eccleston said, ‘“‘and do a spectrum.” Pons also called Robert Hoffman at the radiation safety office on campus. Hoffman had a master’s degree in medical physics and had some

BAD

SCIENCE

@ 87

experience with radiation detectors, but he was not remotely a research physicist. However, Pons trusted him, which may have been more important. Pons told Hoffman that they needed help in observing what Hoffman called “gamma photons.”’ Pons appears to have reasoned that any neutrons emitted by the palladium electrode might be interacting with the hydrogen in the water bath that surrounded the cell and releasing a gamma ray. In fact, the neutrons should have been interacting in just such a manner. This was a well-established nuclear reaction. If Hoffman’s equipment could detect these gamma rays, which always have an energy of 2.22 MeV, it would be as good as seeing the neutrons themselves. And if all the neutrons were disappearing into the water bath, that might explain why Pons and Fleischmann hadn’t seen any. Hoffman was suffering from the flu during the week of March 6, but he offered to do what he could. He carried a portable detector down to Pons’s laboratory and set it up directly over the working cells. The detector was a standard health physics device, known as a sodium iodide detector, and Hoffman connected it to a series of electronic devices—a

photomultiplier tube, a preamplifier, and a device called a pulse height analyzer. This equipment would not only flag any incoming gamma rays but also measure the energy of the gamma ray. If the fusion cells were emitting an excessive number of gamma rays at that one signature energy, 2.2 MeV, it would suggest that neutrons had been emitted from the cell. Hoffman let the detector run for forty-eight hours directly over the fusion cells. Then he measured the background radiation by running the detector for another forty-eight hours over a laboratory sink on the far side of the room. When he was done, he compiled the raw data, which were on pages of computer printout, row after row of numbers corresponding to gamma rays and energies, and handed them over to Pons.”

On March 10, Pons and Fleischmann initiated the process of betrayal. As told by all parties, their actual strategy was not premeditated; however, that it would involve a betrayal was. That day Pons received a call from Ron Fawcett, an electrochemist at UC Davis who was the American editor of the Joumal of Electroanalytical Chemistry. Fleischmann had suggested JEAC as his second choice for publication at the BYU summit conference. They had no guarantee that Nature, their first choice, could accommodate their request for an immediate back-to-back publication, but JEAC was virtually family. Roger Parsons, the managing editor, was a professor at Southampton and had

88 8 GARY

TAUBES

been a close friend of Fleischmann’s since the late 1940s. Fawcett had done his doctorate in Parson’s laboratory at Southampton. Both Pons and Fleischmann had published frequently in JEAC. Fawcett called Pons, as he later told reporters, simply to discuss the performance of one of Pons’s students, who had been out to UC Davis on a job interview. If so, it was a fortuitous coincidence. Pons took the

opportunity to tell Fawcett about the fusion project. He said that they had detected tritium, which he considered proof enough that they were on the nght track. He even told Fawcett about the BYU complications, and his concerns about getting into print as soon as possible. As Fawcett told it, Pons “wanted to get something together real quick’’ and send it to Fawcett by overnight delivery that evening. If Fawcett received the paper the next day, he could stamp it with a March 11 reception date, which could be important in any future priority fights. Fawcett could then route it on to Parsons at Southampton, which is exactly what he did. Fawcett thought cold fusion sounded like a brilliant idea. Typical of Pons and Fleischmann’s genius to come up with something so simple. Why hadn’t anyone else thought of it? Jones and the BYU contingent later viewed the JEAC paper as an out-and-out attempt to shaft BYU and break the universities’ agreement. It’s also possible that it represented afail safe for Pons and Fleischmann. Publishing in JEAC certainly violated the agreement with BYU, but it ensured that they wouldn’t be victimized by fate any more than they already had been. Pons and Fleischmann feared that if they submitted to a refereed journal, as Jones was insisting, one paper—the BYU paper, for instance—could be accepted and the other—theirs—rejected. This was one of the many worst-case scenarios of the back-to-back submission agreement. Their anxiety was understandable. How could Pons and Fleischmann write a paper in the short time available, with the paucity of data they had,"° that would meet the standards of the referees of Physical Review Letters, all of whom, coincidentally, might be physicists? On March 13, an obviously frustrated Fleischmann sent a fax to Dave Williams at Harwell. Fleischmann explained their situation and included a provisional tabulation of the heat data they’d taken so far. He declared that this “information is still very incomplete so you can imagine how annoyed we are to be rushed into premature publication.” Fleischmann added that the rate of tritium generation in their cells was much less than it should be, considering the copious heat production, and neutron production was lower still. “What on earth is going on?”

BAD

SCIENCE

@ 89

he wrote. All of this should explain, he said, ‘““why I am so nervous.” He

closed, “yours exhaustedly.”’" That same day Fleischmann shipped two complete fusion cells to Harwell to run for neutrons. The following day Williams received the two cells and relayed a succinct message to Fleischmann: ‘‘Received cells, all well.’’ This was not exactly the case. Finlay and Sené, who were operating the neutron detectors, hadn’t observed any neutrons from the cells that had now been charging for three days. Also on March 13, Ryszard Gajewski at the Department of Energy called Steve Jones and told him that he had spoken with Stan Pons. Pons apparently said he was still committed to publishing back-to-back articles. But Pons was also still insisting that Jones’s work, as Jones wrote it in his logbook, “follows from his proposal.”’ That evening Pons and Gajewski spoke again, and once again Gajewski relayed the conversation to Jones. This was the same story. Pons was troubled by the fact that Jones would report electrochemical data in his paper, and he repeated the implication that Jones’s work was “‘steered”’ by the proposal. As Jones recalled it, ‘“Gajewski was concerned about this. He really was. Gajewski is very upright. Integrity is the first thing. Gajewski challenged Pons. You want an investigation, fine, we’ll have an investigation. Gajewski was talking to me: ‘What do we do? I’ve been funding you; I need to know about these allegations.’ ” Jones, too, was upset and considered himself very upright. He wanted immediate action. He disliked the idea that anyone was spreading these accusations, and he disliked the feeling it engendered in him: “‘I just start feeling jittery,” he said, “‘like hot pins and needles inside. I don’t like it. I mean, here’s a guy spreading rumors behind my back. I called Stan andI said, ‘Look, I hear again this rumor. I thought we worked through this. I showed you my history. What’s going on? And why are you doing it behind my back, and why aren’t you asking me if you have questions? Why didn’t you raise the questions at the meeting? I’m sorry, but I’m feeling very upset.’ And Stan said, “Well, I’m feeling very upset.’ And he hung up on me.” This was only the beginning. Pons and Fleischmann then called Jones back. ‘‘They had either two extensions or a speakerphone,” Jones recalled, ‘‘and Fleischmann said, ‘Here’s a warning: you could get heat to the degree that things could be dangerous,’ and he told me certain conditions should be avoided. He said, ‘Look, be careful down there

because you could get some real problems.’ ”’ Jones didn’t know whether this was a threat or a warning.

90 8 GARY

TAUBES

In any case, both Pons and Fleischmann now seemed as angry at Gajewski as they were at Jones. They had yet to receive any funding from Gajewski, but they had heard that the DOE had finally approved their proposal on March 2. The two chemists believed that Gajewski was holding the money until the situation between Utah and BYU could be resolved, which meant, as far as they were concerned, that they were being forced into this collaborative arrangement.’? All Jones recalled was that the money had been approved but that Pons and Fleischmann didn’t have it. ““They were peeved about that,” Jones said. “All of this was somehow mixed in there. So I said, ‘Look, guys, all I’m asking is don’t spread insinuations behind my back. If you have any questions ask me.’ And Fleischmann agreed.” Jones then told Jae Ballif what had been passing between the two universities and the DOE. Ballif apparently suggested that the University of Utah should be asked to make a public retraction. And Jones addressed a letter to Ballif outlining his position: . .. I do not believe that I have incorporated any of their original ideas into my research. Rather, I have invited Pons and Fleischman [sic] to make use of our neutron detection equipment, developed at BYU over the past few years for this and other research, in conjunction with a joint research effort. Their methods are complementary to ours and there is therefore good reason to join forces, especially since cold nuclear fusion has potential (though with low probability in my mind) to be of great benefit to mankind. The charges of improprieties are serious and I wish to demonstrate that they are unfounded to an impartial review board, to put an end to these unsubstantiated rumors, and, hopefully, to redeem a severely strained agreement to perform collaborative research on cold nuclear fusion. Therefore I propose that a thorough investigation be conducted. . . .?

On the night of March 15, Pons enlisted legal help of a sort. C. Gary Triggs, his lawyer and childhood friend, had come to town for a strategy meeting that Chase Peterson had called for March 16. Triggs was a bulldog of a small-town defense attorney, fast talking and full of southern homilies. As Peter Dehlinger put it, you could imagine him in court charming the socks off of juries with his yarns and his mellow accent. More than anyone, maybe even Fleischmann, Pons now turned to Triggs for advice. Once again, Gajewski relayed to Jones the details of a caustic conversation with Pons and Fleischmann. As Jones noted it in his log, Pons went

BAD

SCIENCE

8 91

into a “‘tirade”’ and read a prepared statement, alleging, once again, that there had been improprieties in the reviewing process and that Jones had taken information from the Utah proposal and used it in his research. Pons said they would not retract any accusations they had made, and he would not allow himselfto be coerced by Jones or anyone else. Gajewski warned that if it came to an investigation, both universities “would be dragged through muck.” The way Pons and Triggs told it the next day, they had let Gajewski have it with “all their legal guns blazing.’’ The two said that Gajewski had been making veiled threats, to the effect that Pons would never get a DOE grant again if he didn’t shape up. So Triggs, who was apparently listening in on the extension (or speakerphone), signaled Pons, who told Gajewski he would call him back. Then Pons and Triggs hung up and discussed the situation. Triggs wrote up the statement, and Pons called Gajewski back and, with a recorder taping the conversation, read it to

him. According to Hugo Rossi, dean of the College of Arts and Sciences, Pons later said that Gajewski made incriminating statements on that tape. But another witness to this story insisted that ‘‘Gajewski smelled a rat and wouldn’t say a thing.”

CHAPTER MARCH

16TO THE

MARCH

6 23,

1989:

WAGER

I seen my opportunities, and I took ’em. GEORGE WASHINGTON PLUNKITT of Tammany Hall

J f cold fusion had a patron saint, that dubious honor would probably go to Blaise Pascal, the renowned seventeenth-century physicist, mathematician, and philosopher. Pascal renounced a life of science for one of faith, which many of the proponents of cold fusion seem to have done, and he wrote down the terms of the wager that, for him, made

this

choice inevitable. Pascal argued that to bet on the existence of God and to be wrong is to lose little or nothing. To wager correctly that there is a God is to be rewarded with an “‘infinity of infinitely happy life.” “Let us assess the two cases,” he wrote: “if you win you win everything, if you lose you lose nothing. Do not hesitate then; wager that he does exist.” Throughout the cold fusion episode, the proponents of cold fusion would subscribe to the logic of Pascal’s wager. To bet that cold fusion existed and to win was to be rewarded with a payoff that, while not literally infinite, certainly seemed like it at times. To bet wrongly cost relatively nothing: a few million dollars, a few months of work, or a 72

BAD

SCIENCE

@ 93

reputation would always seem inconsequential in comparison to the potential reward.! One year later, for instance, Chase Peterson insisted that he had never

believed that cold fusion necessarily was real, but that what was important was that it could have been real. Here was Pascal’s wager. Peterson said, ““You get burned if cold fusion doesn’t work, but you sure get burned if you don’t do anything about it and it does work. So you’ve just got to be smart.” After March

23, the conventional wisdom was that somehow

the

lawyers had forced the press conference on Pons and Fleischmann. Nobody knew exactly which lawyers, maybe the patent lawyers, but that was not the point. What was certain was that the two chemists would not willingly have initiated such a piece of shameless grandstanding, and the administration of a respectable university surely could not have been responsible for it. So lawyers were the natural scapegoats. But in the final analysis they were only accountable for one third of the responsibility. When Peterson realized that the equation he had worked out with BYU was not a workable one and that cold fusion, with its prodigious potential, could not be left in the hands of the principal investigators alone, he repeated the procedure he had employed with the artificial heart six and a half years earlier. At that time, Peterson had gathered all the key players: the surgeon, Bill DeVries, and his assistant; the regular and intensive-care nurses; the

research team; the lawyers; the hospital security people; the public relations staff; and the hospital administrators who would finance the procedure. Then he said, “I’m not in charge of this, but I’m going to moderate. You describe to me what you think your role will be if we ever choose a patient for the artificial heart.” They went around the room, and everyone spoke honestly and directly. Once they had worked out all the complications to everyone’s satisfaction—and once they had located a viable patient, the redoubtable Barney Clark—they went for it. With cold fusion, Peterson had a considerably smaller team to assemble. He brought in the lawyers: Peter Dehlinger from Palo Alto and two whom Pons had requested personally, C. Gary Triggs and Gary Sawyer, another North Carolina patent expert who had worked with both Pons and Triggs in the past. From the university were Jim Brophy, the vice president for research; Norm Brown, head of the Office of Technology Transfer; and, of course, Pons and Fleischmann.

This meeting convened in Peterson’s office on March 16, and the lawyers began with their agenda. As Peterson recalled, “They began to

94 8 GARY

TAUBES

say, “This is crazy. This has got to be identified now as the PonsFleischmann phenomenon. It’s got to be so identified. And the patent primacy rests partly on lab books and all the rest just on being ahead of the game.’ ”’ Pons and Fleischmann demonstrated to the lawyers that they could document their provenance, which was consideration number one. They discussed dates of invention and decided, on whatever evidence,

which nobody would later admit to actually having seen, that they had priority. Peterson then shifted the discussion to the issue of economic development. He was out to assure that any economic benefits accrued by the technology of cold fusion would go to the state and the university. He had already arranged with Pons and Fleischmann that royalties would be split, one third to the two chemists, one third to the university, and one third, up to a preset limit, to the chemistry department. Now he wanted to assure that there would be royalties, that if cold fusion blossomed into a viable energy source, it would happen in Utah and not Silicon Valley or Japan. Pons and Fleischmann now explained the status of the BYU cold fusion research. Upon reading their proposal, Steve Jones had obviously made a major shift in the emphasis of his research. They believed that his electrolysis experiments from 1986 were just one of many approaches he had tried, and that they hadn’t panned out. They considered meaningless Jones’s notarized lab book page—the one on which he had set down his ideas of cold fusion. They were idle scribblings. Yes, Jones had mentioned palladium, but not in any specific context, and he had never gotten around to using palladium electrodes, or at least not until after reading their proposal, which coincided with his return to the electrolysis experiments. The two chemists said they believed that Ryszard Gajewski was responsible for initiating the dispute, and thus their present troubles. The way they had come to see it, Gajewski had been funding Jones to the tune of a few million dollars over six years and had nothing to show for it. Suddenly they appeared witha startling idea of great promise, one that would certainly justify substantial federal funding if it panned out. And, equally suddenly, Gajewski began to stall. Their funding seemed to be contingent on a collaboration with Jones. They believed that Gajewski was trying to validate somehow all that muon-catalyzed fusion money by linking it to their cold fusion. They had used their connections at the Department of Energy to check Jones’s funding history, and what they

BAD

SCIENCE

# 95

had found was all muon-catalyzed fusion. They had looked up his various publications and found only that one insignificant paper on piezonuclear fusion. Pons and Fleischmann’s most important conclusion was that, like Peterson, they could not trust Jones’s ‘“‘naiveté.”” (“‘Naiveté,’’ Peterson

later remarked, “‘is one way of saying it. There are other ways of saying itzs) Peterson’s congregation then decided that they owed Jones and BYU nothing. They believed that they had developed their technology independent of Jones, and they could let history, the peer review process, or the patent board determine whether Jones had taken anything from them. Norm Brown described it as a case of apples and oranges. ‘““We could let Jones do his thing,” Brown said, “‘and we would do ours, as long as we did ours first. Then Jones could say, “Yeah, me too.’ Or he could say, ‘Well, I found out this other interesting thing,’ which is what he ended up doing. It could hurt us, on the other hand, if we say, “Yes,

Jones invented the same thing,’ which is what is implied by having back-to-back publications. So we ended up thinking that we would be legitimizing Jones, building a mountain out of a molehill, if we did that.” Peterson apparently did suggest momentarily that he call Jeffrey Holland at BYU just to inform him that they were not going to abide by their agreement, but he was quickly convinced that that was not a wise idea. If he gave Holland five days’ notice, or even two days’, BYU might go public immediately. That was a risk they wouldn’t take, so there would be no communication with BYU. As long as the Utah patents were on file—Dehlinger had filed one on March 13 and would have a second filed by the twenty-first—they would retain domestic rights to cold fusion even ifJones announced first. In the United States, patent rights go to the first to invent, which is to say the person who can establish the earliest date for a working prototype of the device in question. In the event of a public disclosure, anyone believing he or she hasa claim to the patent has one year from that moment to file. In all foreign nations, however, with the exception of Australia and Canada, the patent goes to the first to file. Once the invention is disclosed publicly, no patents will be awarded. Thus, any hold on the foreign rights would be lost the instant Jones made his disclosure. Public disclosure has a relatively wide definition: writing up the invention in a company newsletter, for instance; discussing the technology

96 8 GARY

TAUBES

with anyone not party to a secrecy agreement; or releasing any information that would allow an expert to reproduce the invention might constitute public disclosure. The abstract that Jones had submitted to the American Physical Society, even with two meager sentences on cold fusion, might constitute a public disclosure once it appeared in print. It was scheduled for the first week of April. That became the deadline for whatever had to be done,

which unveil second appear

gave the U roughly three weeks. The option of having Pons cold fusion at the American Chemical Society meeting in the week of April was no longer viable. Jones’s APS abstract would a week before the meeting.

Peterson,

Pons,

and Fleischmann

had bandied

about

the idea of

throwing a press conference since the BYU summit meeting. The two chemists had already discussed the situation with Roger Parsons and Ron Faweett, the editors of the Journal of Electroanalytical Chemistry, and had received their assurance that a press conference would not jeopardize the publication of their paper. (In fact, JEAC’s decision to publish was somehow treated as justification for the announcement. That the paper would not be peer-reviewed, but rather accepted on the strength of Parsons’s judgment alone, was considered atrivial technicality.)? Initially it was Peterson who suggested the public announcement, but the three lawyers apparently embraced its wisdom. Dehlinger later supposed that the decision might have gone the same way even if everyone but Peterson had been against a press conference. But such was not the case. ““The fact is,” Dehlinger said, “‘the three lawyers were arguing that there is no second place in this kind of business. Either you’re there first or no one remembers you. Gary Triggs was particularly adamant because he is such a fanatic believer in Stan that he was the most upset at the idea of sharing this with Steve Jones.”’ Dehlinger concluded, “We all, I think, supported and were trying to overcome the resistance of Stan and Martin.”’ Pons seemed most concerned with the logistics problem at hand. How would they write the Nature paper, finish the JEAC paper, prepare a press release, and complete the necessary research, all in one week? But he agreed to go along with a public announcement. He trusted Triggs, perhaps even as much as he trusted Fleischmann. As Dehlinger recalled it, “Gary Triggs never doubted the science. And with that position he also never doubted that this was one of the great scientific breakthroughs of the twentieth century, and his friend Pons was going to get credit only if he was bold, and he was certainly pushing Pons to be as bold as possible.’” Norm Brown later observed that the relationship between

BAD

SCIENCE

8 97

Pons and Triggs transcended that of lawyer and client. ‘‘Stan relied on Tnggs a lot,’’ Brown said. “Pons looks to him for advice, and so, de

facto, if Pons wasn’t satisfied, the university had a problem, and if Triggs wasn’t satisfied, Pons wasn’t satisfied.”

Dehlinger’s recollection of the meeting also had Fleischmann “‘almost in tears’ as the consensus finally emerged that they would call a press conference.

This contradicts Peterson, who later said he would never

have gone on with the announcement had Fleischmann been so noticeably against it. Maybe so. Either way, Fleischmann certainly was the most prescient about the ugliness of the deluge that would follow a news conference. And afterward it was Fleischmann who would lay the entire responsibility for the decision and the subsequent circus on the U administration. ““That was the decision of the university,” he said. ““You can read into that anything you want.” Nonetheless, at that point, Pons and Fleischmann still could

have put a stop to the affair. They did not. The administrators and lawyers had only to worry about whether Pons and Fleischmann were correct in their interpretation of their experiment. Peterson later said that he believed in cold fusion, to the extent that he

did, because Pons and Fleischmann were “‘very competent electrochemists. They say that they’ve got something that cannot be explained by a chemical reaction. Well, these guys ought to know what a chemical reaction is.’”’ He also believed because Jim Brophy believed it unconditionally, and Brophy had been trained as a physicist. And Peterson was impressed with Fleischmann’s brilliance, as was everyone else. Although Fleischmann was not a physicist, he appeared to have a deep understanding of the field. It was the account of the explosion, however, the famous meltdown,

more than anything that convinced the lawyers and administrators. All the other evidence of nuclear fusion—heat, neutrons, gamma rays, and

tritium—paled next to the tangibility of the explosion. “Frankly,” said Dehlinger, the biophysicist turned patent attorney, “if there was one thing that made me feel good and think that there is something there, it was the explosion. Yes, it can be chemical, but that didn’t seem likely.

Something pretty significant in terms of heat generation must have been happening. It was the anchor which many of us were using. Whenever you're doing science, lots of things can go wrong and get in the way of reproducibility. We took comfort in the few events which seemed spectacular.” In a sense, Peterson and his congregation ultimately believed in cold fusion because it was too big to question. It was Pascal’s wager. “If there

98 0 GARY

TAUBES

is any merit to this,’’ Peterson said later, ‘‘can we afford to let it go the way of classical, normative science, that is, a very dignified, reserved

announcement, a very slow and methodical process of development and substantiation?” The answer was no. One final consideration still might have prevented a press conference, however. Brown suggested that before going public they give serious thought to the political implications of their announcement. Pons and Fleischmann, after all, were claiming that it would be possible to build

nuclear weapons with the technology. Brown observed that should this be the case, as with Pandora, there’d be no going back once they opened the box. ‘““What if this gives Qaddafi the ability to make a nuclear bomb for fifty dollars?’’ asked Brown. “Is this something that we really want to do without thinking about it ahead of time?” This was somehow too profound to contemplate in the short time they had, so it was ignored. Brown recalled that Peterson and the others gave it a few seconds’ thought and in effect said, “Okay, now we’ve thought about it, let’s throw a press conference.”’ As the meeting concluded, Pam Fogle, the university’s news director, was called in to discuss how to handle the press. She was informed that her office could interview Pons and Fleischmann the following day, which was Friday the seventeenth; then she would have the weekend to write a press release. It would have to be ready by Monday so that the scientists, lawyers, and administrators could review the draft. Be careful,

Fogle was told, that no drafts of the press release fall “into the wrong hands.”’ It sounded like a spy novel. Fogle said the determination was made then to schedule the public announcement tentatively for March 23, one week later. Once they had committed themselves, secrecy became a greater concern than the validity of the science. Peterson, for instance, chose not to consult any of the

physicists on campus. He feared they might leak news of the ““F-word” to the outside world, or even leak word of the press conference to Jones. He also believed his chemistry department was the “more prestigious department.” Thus, employing logic that seems dangerously specious, Peterson concluded that his prestigious chemists did not need assistance from physicists to solve a physics problem. Peterson did seek advice from Hans Bethe, one of the legendary figures in nuclear physics. Bethe, who was eighty-two years old, still taught quantum physics at Cornell. Physicists liked to say that taking quantum mechanics from Bethe was like taking Russian literature from Tolstoy. (Peterson and Bethe were distantly related through their children’s marriages. Bethe’s daughter-in-law had a brother who was mar-

BAD SCIENCE

#99

nied to Peterson’s daughter.) Peterson phoned Bethe and told him that Pons and Fleischmann had observed electrochemically induced cold fusion. Bethe replied that this sounded very unlikely. Then Peterson said that physicists at Bigham Young University were also claiming cold fusion and were going to publish a paper. He didn’t want his university left out. “Let BYU publish alone,” Bethe said. “‘Let them make fools of themselves.” Peterson ignored the advice, apparently because it was not what he wanted to hear. And Bethe wasn’t swayed by the persuasiveness of Pascal’s wager. Indeed, as far as the wager went, to sit quietly and let Jones and BYU publicly announce the discovery of cold fusion was, in effect, to bet that cold fusion did not exist. Bethe, with his deep under-

standing of canonical nuclear physics, might be able to do that. Peterson could not.’ Peterson spent March 20 in Washington on business unrelated to cold fusion. He flew back to Utah the next day. On the flight were two Utah physicists, Pierre Sokolsky and Michael Salamon, who had been visiting their benefactors at the Department of Energy. Half of the remaining passengers, or so it seemed, were adolescent girls returning from a Washington field trip. Sokolsky and Salamon escaped their chatter by hiding in the galley. Peterson also fled to the galley. After some brief small talk, Peterson told the two physicists that he’d been doing some reading, and he’d like to know what they could tell him about muon-catalyzed physics. Salamon and Sokolsky were astonished. Muon-catalyzed fusion was esoteric even within the refined field of nuclear physics. It was certainly not a subject that they would have expected to pique the interest of a man of Chase Peterson’s medical and administrative background. In fact, the two physicists had never given it much thought themselves, although as physicists they knew of it and could explain it, which they did. For the better part of an hour, Peterson interrogated the two on yarious fusion mechanisms, using correctly all the technical jargon. Later Salamon and Sokolsky recalled that Peterson never explained the motivation for his curiosity. Nonetheless, their opinion of Peterson, which

had not been high, skyrocketed. On March 21, Dave Williams at Harwell faxed a succinct message to Martin Fleischmann in Utah:

100 8 GARY

TAUBES

5 days now & no neutrons. Any suggestions?

Fleischmann called Williams and told him they could not sit on cold fusion much longer. Now, however, Fleischmann was optimistic. He

said he was confident that he had counted neutrons, having detected apparently three times as many coming from the vicinity of the fusion cells as from the background. Although this was not the billionfold or trillionfold increase predicted by nuclear physics, it sounded convincing to Williams. (But Williams later admitted, “I don’t know anything about neutron counting.’’) Fleischmann also said that they had been detecting gamma rays from a cell with an eight-millimeter palladium electrode but that these rays had vanished. They did, however, seem to be registering a very definite gamma ray signal from a cell with a four-millimeter electrode. ‘“‘And so he concluded,” Williams recalled, ‘‘that the eight-millimeter rod had died. Therefore there was a possibility that the experiment could die, or maybe not even work. So he knew at that time that the thing was irreproducible.”’ The effect may simply have required a fine tuning of the cells that was hit-or-miss. Some cells worked and some didn’t. Williams had his fiveday-old cells removed from the neutron counter at Harwell and three new cells inserted. That same day, Pons telephoned Steve Jones, wanting to know if he still planned to publish his paper.* Jones said it was almost written, and they were committed because of the APS meeting. The two then agreed that they would rendezvous to send off the papers at the Federal Express office at the Salt Lake City Airport, at 2:00 p.m. on March 24. Fleischmann was returning to England that day, and it would also be the day after the Utah press conference. Jones noted in his lab book that when he asked Pons if the Utah cold fusion paper might be submitted early, “e.g. by Fleischman [sic], he said no.”’ So Jones was suspicious. He simply had not asked the right question.

On the morning of March 22, Pam Fogle began alerting the local and national press to the coming news conference. Once she called Jerry Bishop of The Wall Street Journal, the story began to leak, as the administration had feared. Bishop was a ponytailed Texan whose working wardrobe ran to cowboy boots and blue jeans and who did not look like a reporter for the country’s preeminent conservative business paper. But he had been a science reporter for three decades. When Fogle told him

BAD

SCIENCE

@ 101

that Utah had sustained a nuclear fusion reaction at room temperature for one hundred days, Bishop refused to believe her. After talking to Fogle, Bishop went to lunch with a free-lance science writer who was connected with the Council for the Advancement of Science Writing. Bishop described the call, and the friend reminded him of a recent meeting at which a physicist from Utah had discussed some kind of strange fusion. Bishop went back to his office, checked through

his files, and found a copy of a Steve Jones paper on muon-catalyzed fusion. Bishop called BYU, managed to reach Paul Palmer, and asked what the story was with the University of Utah announcement. “All we can say is our results don’t confirm theirs,’ Palmer told him. From that moment, Bishop knew he had a story. His story in the Joumal the next morning reported that a new fusion breakthrough would be unveiled and speculated that it would show “that hydrogen atoms can be forced to fuse together inside of a solid material.” Bishop reviewed Jones’s previous work in muon-catalyzed fusion and noted that both the University of Utah and BYU had simultaneously submitted papers to

Nature.* As it turned out, the Financial Times of London beat Bishop to the

story. It seems Fleischmann had gone to an old friend at Southampton, Richard Cookson, a retired chemistry professor, to get advice on giving the story to the British press. Cookson put Fleischmann in touch with his son Clive, who was a reporter at the Financial Times. Young Cookson had explained to Fleischmann that the Financial Times would not be published on March 24 because it was Good Friday. If Fleischmann wanted the British press to get the story before the long weekend, he would have to let them publish it on Thursday, the day of the press conference. Fleischmann apparently talked it over with Pons and maybe Brophy, then gave his permission and the necessary informa-

tion. Harwell physicist Ron Bullough suggested that Fleischmann’s act was motivated by national concern. “‘Since he’s a true Brit,” said Bullough, ‘the felt he had to do that. I think he was using Cookson to warn the establishment and the government that all hell was about to break loose.””® With Cookson and Bishop working the story on Wednesday morning, the news immediately got back to BYU. Like all rumors, the story appears to have evolved spontaneously as it spread. Bishop apparently * Actually, that had not yet happened and was looking less and less likely.

102 8 GARY

TAUBES

tracked Jones back to Gajewski, whom he called to find out what he knew about the Utah press conference. That prompted Gajewski to call Jones and tell him that “‘all hell has broken loose” at the Department of Energy.

As Gajewski heard it, apparently misunderstanding Bishop, the University of Utah had a press release that claimed heat production by cold fusion, while simultaneously claiming that a reviewer of the proposal confirmed the result. Jones, of course, was outraged. ‘‘Baloney!’’ he scrawled in his lab book:

And why should they announce our unpublished results to the press? Press release also flies in face of our agreements not to speak of results publicly until papers back-to-back were in (Friday 3/24/89).

Before the day was out, Jones also received calls from the Financial Times, to confirm that he had confirmed Utah, and from a broker in

Boston, who apparently wanted to know if they really used palladium, in which case should he buy futures? As the reports of BYU’s confirmation escalated, Grant Mason, dean of

the College of Physical and Mathematical Sciences at BYU left a message for Paul Richards, the university’s public communications director: Apparently the U of U has released a story to the press indicating that BYU backs up or supports some research they are doing on cold fusion. [Mason] is concerned that this could be quite sensational since we do not in fact support their conclusions. We need to know if the press release went out and exactly what it said, and he also wants to prepare a statement as a disclaimer.

Richards called Ray Haeckel, his opposite number at the University of Utah, and Haeckel read him a copy of the press release, which made no mention of any confirmation whatsoever. “I can’t imagine how you have this,” Haeckel said, ‘“‘because we haven’t even sent it out yet. We’re not going to send it out until tonight.” Richards reported back to Mason, who promptly phoned Jim Brophy. Brophy also insisted that Utah hadn’t issued any press release, which was still true. They had been working on a draft, Brophy said, and they had been very careful not to mention BYU at all. “But people are quoting us,”” Mason said. ‘“They’ve got something.” “If you find out where it came from,” said Brophy, “let me know. Because I don’t know where.”

BAD

SCIENCE

#8 103

Mason then warned Brophy that should Utah hold a press conference, the BYU administration would consider it ‘‘a stab in the back.” The phone calls continued through the afternoon. Steve Jones called Joe Ballif with the news that Pons and Fleischmann might be holding a press conference. Ballif refused to believe it: a Utah press conference constituted betrayal of a magnitude he simply could not accept. He had Chase Peterson pulled out of a meeting to take his call. He said that he’d heard a rumor of a press conference, which must not be true. And Peterson replied, ““We have talked about that, yes. We intended to make some kind of an announcement.” Peterson went on to say that he’d never been satisfied with the outcome of the meeting in Provo. He said he’d call Ballif back, which he never did.

The announcement was made the following day in the main foyer of the Henry B. Eyring Chemistry Building on the Utah campus. The available seating space was quickly taken by press, administrators, students, and curious scientists. Marvin Hawkins attended the press conference with his wife, his father, and his in-laws, and managed to find seats three rows back. Fleischmann had warned him that their lives would never be the same after they went public and that the activity unleashed would be astounding. Hawkins thought he was prepared for the worst. He would say later that he knew that something ugly was going to happen. Still, he added, ‘in my wildest imaginations, I never expected what happened.”

CHAPTER MARCH

23

AND

7 24,

1989:

AFTERTHOUGHT

6, two members of the BYU administration attended the Utah press conference. One was Julie Walker, the electronic media expert in the public affairs department. The other was Alan Knight, an advertising sales representative for the BYU alumni magazine, BYU Today. Knight happened to be in Salt Lake City that morning, so he volunteered to help Walker make a tape of the announcement. After the announcement, Knight went to pick up a copy of the press release and found himself face-to-face with Chase Peterson. Peterson said, ““Gee, I’m glad somebody from BYU came up. We invited BYU to be a part of the press conference. They turned us down.” When Knight returned to BYU with this tale, it ‘raised some eye-

brows,” in the words of Paul Richards, the PR director. One assumes

that Walker’s tape of the conference raised a few more. At one point in the proceedings, a reporter asked whether Pons and Fleischmann were aware of any similar work going on elsewhere. “Let’s see, I’ll answer ity, 104

BAD

SCIENCE

@ 105

Brophy said. ‘““We’re not aware of any such experiments going on. There are none reported in the literature.” Down in Provo, Jae Ballif and Jeffrey Holland spoke to a gathering of the BYU Faculty of Science. The two administrators explained that agreements had been made with the University of Utah and apparently been violated and that they were going to do everything they could to protect the integrity of both Steve Jones and the institution. Said Ballif, ‘““We wanted them to know that we were going to act honorably, and that we were going to protect the integrity of our people.” Afterward, Ballif and Holland walked to the physics building, where Jones and his colleagues were working furiously to finish the paper. By this time their various sources had confirmed that Pons and Fleischmann had already submitted a paper to a journal other than Nature. Since this was a flagrant violation of the agreements, they were now justified in submitting their paper to Nature as soon as possible. While Pons and Fleischmann were still doing post—press conference interviews in Pons’s basement laboratory, Bart Czirr faxed off the BYU paper on piezonuclear fusion to the Washington office of Nature. That night Jan Rafelski, who was in town to help, treated his BYU collaborators and their wives to dinner at a local Mexican restaurant. Paul Palmer would recall that it felt like a victory celebration: “We just wondered what our victory was; we decided our victory was that we were now separated from those guys.” This still left open the question of whether to abide by the now many times violated agreement to meet at the airport at two o'clock the following afternoon and send off the simultaneous submissions. Jones said he was tired of the charade. He had planned to spend the Easter weekend in Denver with his family. He stayed home on Friday to pack. Friday morning, Paul Richards called Palmer and said that a local television news team had heard about the airport meeting. Apparently these broadcasters pictured the joint submission as a modern-day equivalent of the golden spike ceremony, which had celebrated the joining of the east and west lines of the transcontinental railroad. The news team wanted to know if BYU was going to go through with the meeting. Czirr and Palmer called Jones at home and asked for his opinion. Jones

suggested Czirr go because he would be hard-nosed about the charade. Czirr was flattered but said he wouldn’t do it. They’d already submitted

the paper. Czirr observed with pungent candor that this was both ‘“‘immoral and stupid.” He said, “Pretend on national television, big event,

106 § GARY

TAUBES

old brotherly love here. They betrayed us Thursday and we’ve already submitted. It’s a big farce. I won’t go.”’ Czirr and Palmer then called Paul Richards, who suggested maybe he would go, but Czirr badgered him out of it. That was the end of it. ; The television station sent a cameraman, who filmed Marvin Hawkins

waiting around the Federal Express office with an envelope in his hands.'

Book

Tue

A COLLECTIVE DERANGEMENT OF MINDS Men, it has been well said, think in herds; it will be seen that

they go mad in herds, while they only recover their senses slowly, and one by one. CHARLES MACKAY, Extraordinary Popular Delusions and the Madness of Crowds, 1841

There is something fascinating about science. One gets such wholesale returns of conjecture out of such a trifling investment of fact. MARK TWAIN, Life on the Mississippi

ty ~~

;

A

a

a

“7

q

a

ia

reo »s

.

.

wir

em cy

Sab

§

oes

Sr 4) %

aais



.

v

a¥i(941109 TWIMIGHARIG | oe.

10

ie

CHM = ato

pe i r



.



_ 5 al

j



!

Ay 7 La HA

eh *i i an STORY odie pa siniaeaaie

oa pli

a ee

a

a

as

oe > a = P «_ ;

Tt

oe

an *, i

|ee

Ds

ed

ore Ve

>

i

1 AUSTIN,

TEXAS

At 4:45 in the afternoon on March 23, Al Bard found a note from his

secretary that Mark Blackwell of the Dallas Times Herald had called about “Pons and Fleischmann fusion.” Bard was a professor of chemistry at the University of Texas in Austin and was considered by many to be the best electrochemist in the country. He was the editor of the Journal of the American Chemical Society, which was considered the premier scientific journal in the field. Bard was a natural source for reporters who were looking for an informed opinion on a chemistry story, and he knew both Stan Pons and Martin Fleischmann well. He went back thirty years with Fleischmann, and when Utah had hired Pons, the chemistry department had asked for a recommendation from Al Bard, who wrote that Pons was a “‘very good grab” for a chemistry department like Utah’s, which was one of the top twenty in the country. But that was then. Pons and Fleischmann fusion? “You really got this screwed up,” Bard said to his secretary after reading the message. Then he phoned Blackwell, who described what he knew of the announcement out of Utah. “Did they see neutrons?” Bard asked. “Did they detect tritium?” Blackwell said he would find out. An hour later he called back. Yes,

Pons and Fleischmann reported neutrons, and tritium; they detected more heat coming out of the cells than could be accounted for by the energy going in. “So,” Blackwell asked, ‘‘what do you know about Pons and Fleisch-

mann? Do you believe this?” “They’re both respectable scientists,” Bard said. “It’s an amazing effect. If it’s true, it’s terrifically important.” This is what scientists invariably tell reporters when referring to seemingly bizarre discoveries of which they have no knowledge. Indeed, Bard didn’t know what else to say. It seemed so strange.

109

110 © GARY

TAUBES 2 COLLEGE

STATION,

PASADENA,

TEXAS,

AND

CALIFORNIA

Chuck Martin was driving home from his laboratory through the miles of student housing that seem to constitute the better part of College Station, Texas. He was listening to National Public Radio’s All Things Considered when the newscaster reported that two Utah chemists claimed the discovery of nuclear fusion in a test tube. Martin all but ran his car off the road. Martin was a rising electrochemist at Texas A&M anda close friend of Stan Pons. He considered Pons not only a very good scientist, maybe even a genius, but a generous and thoughtful man. Just two days earlier, Martin had spoken to Pons, who had told him to watch the news on Thursday, the twenty-third. Martin had forgotten. Now he remembered: “I can’t believe this. This is what Pons was talking about. My God!” Martin tried to call Pons but couldn’t get through. He then tried Nate Lewis in Pasadena. Lewis, at age thirty-three, was two years younger than Martin, but they had come up through the ranks of academia together and had known each other since 1982. Lewis was an electrochemist at the California Institute of Technology and was considered by his colleagues to be a prodigy of a sort. Martin put it this way: “I’m not as young as Nate, and nobody’s as smart as Nate.”’ Unfortunately, Lewis was ona flight back from Washington, D.C. His wife, Carol, told Martin that he wouldn’t be back until late—perhaps ten o'clock. Martin told her to tape the Cable News Network, which was broadcasting excerpts of the Utah press conference. He called back at nine o'clock; Lewis was still en route. Finally Lewis returned, and his

wife relayed the message, along with the editorial opinion that Chuck Martin had gone crazy. Lewis watched the videotape and then called Martin, who was already compiling a list of the equipment required to induce cold nuclear fusion. To Martin, cold fusion represented the opportunity of a lifetime, manna from the heavens. “How many times in your life,” he said, ‘‘do you have the chance to do an experiment that could impact all of society? This is potentially the most important experiment of the twentieth century. And I’ve been an electrochemist for a decade, and all of a sudden electro-

chemistry is the most important issue in science. And I can make a contribution. My God, it has fallen right into our laps.””

BAD

SCIENCE

8 111

“I’m going to go do this,’ Martin said. “I can’t wait. It’s unbelievables? “Well,” said Lewis, who was not so tight with Pons, and thus wanted

to see a published scientific paper before he bought into it, “I’m not going to do it until tomorrow.”

3-CAMBRIDGE,

MASSACHUSETTS

Half a dozen young chemists were drinking at the campus bar at the Massachusetts Institute of Technology when Richard Crooks, who was a postdoctoral fellow, pointed to the television screen, which was, as

usual, turned to the news. ‘““What’s Stanley Pons doing up there?” Crooks asked. These chemists worked for Mark Wrighton, chairman of the MIT chemistry department, and all of them had heard of Stan Pons. They had also finished three or four pitchers of beer already that evening. What followed was a story they would tell many times over the next few years. The pub was loud, and the volume was off on the television, so they started shouting to the bartenders, ““Turn up the volume! Turn up the TV!” But the bartenders couldn’t hear them. So they sat and watched as the images of Pons and Fleischmann flashed on the screen, followed by pictorial diagrams of fusion, two billiard balls coming together, fusing. And they all looked at one another. Pons and Fleischmann? Fusion? What was going on here? One

of them said, ‘Well, if it’s important we’ll find out about it

tomorrow.” Then they returned to their beers.

:

4 THE

PRESS

Friday morning, The Wall Street Journal page one headline was nearly impossible to miss: TAMING H-BOMBS? TWO SCIENTISTS CLAIM BREAK-

THROUGH IN QUEST FOR FUSION ENERGY. IF VERIFIED, UTAH EXPERIMENT

112 © GARY

TAUBES

PROMISES TO POINT THE WAY TO A VAST SOURCE OF POWER. Then below this: BATTERIES AND PALLADIUM WIRE.

SALT LAKE CiTtY—Scientists working at the University of Utah made an unprecedented claim to have achieved a sustained hydrogen fusion reaction, thereby harnessing in the laboratory the fusion power of the hydrogen bomb. The two scientists said that with no more equipment than might be used in a freshman chemistry class, they had triggered a fusion reaction in a test tube that continued for more than 100 hours.

The story mentioned only briefly that the University of Utah had not released a scientific paper with their announcement, which meant that the scientific community had no way to make an informed assessment of the results. Of those who had ventured uninformed opinions, the article only said that many “expressed a gut reaction of incredulity.” From day one, The Wall Street Journal would be first and foremost in cold fusion. It was rumored that the Journal editors felt they had been embarrassed by The New York Times on the room-temperature superconductor story, which had broken

two years earlier. Thus, the editors

elected to win this one at all costs. Within a month, the Journal’s seeming omnipotence would have scientists calling it, not without some displeasure, The Wall Street Journal of Physics or The Wall Street Journal of Cold Fusion." In fact, the Wall Street Journal staff had been so smitten with the cold fusion story that they were surprised Friday morning to find that their competition at The New York Times had run the story not on page 1 but at the bottom of page 12 under the headline NUCLEAR POWER GAIN REPORTED BUT EXPERTS EXPRESS DOUBT. When the Journal editors saw the Times’s treatment, they promptly phoned their bureau in Los Angeles and asked how the Los Angeles Times had played it. They were told the L.A. Times had also run cold fusion on the front page (PAIR PROCLAIM NUCLEAR FUSION BREAKTHROUGH . . . SCIENTISTS IN UTAH SAY SIMPLE TABLE-TOP DEVICE PRODUCES MORE ENERGY THAN IT USES IN TESTS).? “Thank God,” Jerry Bishop said. ‘““We thought we had our necks stuck Way out.” Bishop saw the story as his own. He might have played it down, as the

uptown competition had, if not for the unorthodox method of announcement. After all, this was an affair endorsed and promoted by the president of the University of Utah. “It was kind of secondary whether it was true or not,” Bishop said. “Just the fact that the claim was made

BAD

SCIENCE

#113

under the conditions that they made it. If Pons had called me directly, I would’ve said, Go away.” Meanwhile, The Washington Post ran a cautionary story on page 3: SIMPLE TEST SUSTAINED FUSION, SCIENTISTS SAY. PHYSICISTS DOUBTFUL OF PRACTICAL APPLICATIONS. Philip Hilts, the Post reporter, communicated his personal skepticism in the third paragraph: “‘Other scientists expressed doubt about the work, however, and one physicist familiar with it said

the announcement was ‘blown far out of proportion.’ ”’ What was recurrent in all these first cold fusion stories was the obvious comparison between Pons and Fleischmann’s absurdly simple contraption and the mammoth multimillion-dollar efforts to induce fusion with conventional methods. The Journal quoted an unnamed source at the University of Rochester saying that the conventional laser fusion program there might “‘ignite a hydrogen pellet within three to four years.”’ Cost to the taxpayer of this “inertial confinement” approach to nuclear fusion: $150-160 million a year. The Los Angeles Times noted that the Department of Energy was spending $350 million in research on magnetic confinement fusion, which was the other conventional approach to fusion power. With all this cash, scientists at Princeton’s Plasma Physics Lab “‘were able to generate temperature ten times hotter than the core of the sun, about 360 million degrees Fahrenheit, but they were able to hold it there for only a fraction of a second.”’ Charles Barnes, an astrophysicist from the California Institute of Technology, noted in the L.A. Times that “a few thousand dollars will get you a few quarts of heavy water and a few hundred will get you the palladium and platinum.” He seemed to be saying, What are we waiting for? After the Utah press conference, the reporters had been invited to Pons’s basement lab to view the fusion reactors. These were situated in a back room

of the four-room

laboratory,

sandwiched

unheroically

between a janitor’s closet and a utility room. The apparatus was less than awe-inspiring, but, as the press release had said, about what one would expect from a freshman chemistry course. A stainless steel water bath, the size of a large picnic cooler, sat on a table in the center of the room. Inside that were two fusion cells, sophisticated test tubes, submerged

neck-deep in the water, wires emerging to connect the electrodes inside to a stack of power supplies. Another three fusion cells sat off to one side, bubbling away like freshly poured soda water. That was it. All five fusion reactors and their accompanying electronics could have been loaded

114 BP GARY

TAUBES

comfortably into the back of a station wagon. Two car batteries were even sitting in plain view behind the apparatus, giving the distinct impression that the power supplies for room-temperature fusion could be purchased at Sears or Pep Boys.

5 SALT AND

LAKE

POINTS

CITY,

PROVO,

SOUTH,

UTAH

At the public relations office at the U, a kind of euphoric madness reigned. The office received over 400 phone calls in the first days of cold fusion (or over 400 phone calls an hour, depending on the source) from, among others, major corporations, investors, management consultants, booking agents, well-wishers—‘‘Congratulations on your incredible discovery. We are very happy to be alive during this time in humanity’s history, and we are looking forward to experiencing a world free of pollution, etc. etc.’’—old friends, new friends, senators, congressmen, foreign consuls, publishers hoping to sign Pons for his memoirs, filmmakers, not to mention Texans fearing the discovery would destroy their already shaken oil-based economy, and, of course, scientists and report-

ers. Friday morning, Chase Peterson attended a scheduled meeting of the University of Utah’s board of regents. The timing was serendipitous to say the least. Peterson gave a brief report on cold fusion, at which point Ian Cumming asked, ‘“What’s next?” [an Cumming may have been the state’s single most influential citizen. He was a community benefactor who shunned the press, sat on boards, and gave pep talks to luncheon clubs on the Utah economy. His personal point of view was “Utah must do more—spend more if necessary—to encourage job development in the state by attracting industry from without and cultivating industry within.’’ It was said around town that Cumming would be the man to choose the next governor of Utah. “What do you mean what’s next?’ Peterson said. ‘“What’s the next step?”” Cumming asked. “Challenge and confirmation,” said Peterson, ‘“‘and we’ll know in six

months or a year what goes on here.” “We can’t wait that long,” Cumming insisted. ‘“We can’t let this thing

BAD

SCIENCE

#115

get out of our hands. We’ve got to keep centrality to it in Utah. And so we need to get some money from the state to get this thing moving promptly.” They immediately phoned the governor, Norman Bangerter, who was relaxing at his vacation home in southern Utah. Cumming then had his private jet pick up the regents in Provo, and they all flew south to meet with Bangerter. It was an impulsive gesture, which, in the words of Hugo Rossi, “would torque the whole affair up several significant notches.” The regents, led by Cumming and Peterson, unfolded this great discovery for the governor. After they discussed it for twenty minutes, he said, “Okay, you didn’t come all this way just to tell me about it. How much money do you want?’ Bangerter’s guests looked momentarily startled. Then one of them suggested $2 or $3 million might do it. Bangerter offered $5 million, “‘if this is as good as you say.” This seemed to be the governor’s attitude about cold fusion throughout: “Knowing nothing about it,” he would say “I am highly optimisHee Under the circumstances, Bangerter’s $5 million offer may be considered cautious. Miles Crenshaw, for instance, a local personality and radio

talk-show host, suggested the state divert $300 million to cold fusion research, with $1 million tax-free to Pons and Fleischmann as an expression of gratitude. Another state senator quickly suggested that citizens be able to invest a portion of their taxes in cold fusion research, to be paid

back at a later date through the proceeds reaped by the state. It was a heady time.

6 AUSTIN

“Friday,” Al Bard said, ‘‘all hell broke loose.” Bard had lived in the heart of Texas for three decades, but his native

Bronx accent hung on stubbornly. He had gone to Bronx High School of Science, City College of New York, and then Harvard, where he first studied inorganic chemistry, which was then a “‘hot”’ field. He studied the discipline under a young British assistant professor named Geoffrey Wilkinson, and even a graduate student could tell he was doing hot

116 9 GARY

TAUBES

work. Harvard, however, announced that Wilkinson would not be given

tenure, and Wilkinson left shortly thereafter. (In 1973, Wilkinson shared the Nobel Prize, proving that Harvard could be as fallible as any other university.) Bard, meanwhile, had to find a new mentor. He got “a little bit interested”’ in electrochemistry, and James J. Lingane signed him up. Lingane had been the student of I. M. ‘‘Pete”’ Kolthoff, who was the father of electrochemistry in America, at the University of Minnesota. Under Lingane, Bard learned to use electrochemical techniques as tools to attack other interesting chemical problems. This made Bard’s reputation, and, with his peers, it launched a renaissance in electrochemistry,

which happened to coincide with the phasing out of those departments at schools like Harvard and Princeton. This explained, among other things, why Bard went on to Texas. Bard spent much of Friday, March 24, trying to glean what facts the news reports provided on cold fusion, which he then recorded in his lab book. He learned that Pons and Fleischmann had worked on cold fusion for five years, that it used equipment along the lines of that in a freshman chemistry experiment, that the sophisticated test tube was the size of a drinking glass; that it used 99.5 percent heavy water, that the reaction took ten hours to get going and continued for more than a hundred hours, past the point at which the power consumed was greater than the power generated—what physicists would refer to as break even—and that it produced four watts of power. The Wall Street Journal and the Dallas Times Herald both confirmed that Pons and Fleischmann had detected neutrons and tritium. According to the Austin American-Statesman, any electrochemist could reproduce the experiment in an afternoon. It looked, Bard thought, like a snap. Bard then wrote down questions to which he wanted answers: for instance, how could there be sufficient pressure within the palladium to cause fusion? A deuterium nucleus is composed of a single proton and a single neutron, so the nucleus has an overall positive charge. That means two deuterium nuclei would naturally repel each other. In order to induce fusion, this repulsive force has to be circumvented or overwhelmed, which can be done with sufficient pressure and heat, as in the cores of stars. In the 15-million-degree plasma at the core of the sun, for instance, the nuclei are moving so fast that they smash into one another and fuse without hesitation. But how could this be achieved chemically? How could a palladium rod and a small electric current force the two deuterium nuclei to come together? It didn’t make sense. At noon Bard met with his research group to tell them what he knew, and that he was going to do cold fusion experiments: “‘It’s not that we’re

BAD

SCIENCE

8 117

going to make any big contribution now,” he said. “It’s done. Ifit’s true, it’s true, but it’s important to reproduce this result.’’ Indeed, Bard felt as

he had when he was young and had read about aclassic scientific experiment and wanted to reproduce it. Not because he wanted to write a paper about it but because he wanted to do that experiment with his own two hands. “I think what most of you are doing is important enough,’ Bard said to his researchers. “I don’t think you guys ought to jump off and get into this unless you really feel very strongly motivated about it. And besides that, this experiment is potentially very dangerous, as far as I can figure out. So the lab where I’m doing this is going to be off limits.”’ And Bard added, “I don’t want you to feel that I’m doing this all for myself. If you guys want to do it, you’re welcome to.” The graduate and postdoctoral students seemed to think that cold fusion, even with all the hullabaloo, was improbable at best. They didn’t

want to get involved. One postdoc, Norman Schmidt, had just joined the group, and he volunteered. So it was Bard and Schmidt alone. That Bard, in his late fifties, actually ran the experiments himself was unusual. Most chemists of his age and caliber consider themselves managers of a sort and believe that the laboratory is a place for graduate students and postdocs. So did Bard, until cold fusion. Bard tried to call Stan Pons, but he couldn’t get through. At 2:00 p.M., Bard received a call from Larry Faulkner, who was dean

of the College of Arts and Sciences at the University of Illinois and had studied electrochemistry under Bard. Faulkner said that he didn’t believe the cold fusion announcement. At 5:00 p.m. Bard heard from Mark Wrighton of MIT, who said he was withholding judgment, but the MIT physicists, on the whole, were skeptical. Wrighton said that they were worried about neutron radiation. If Pons and Fleischmann were producing four watts of fusion power, then where was the radiation that not only is a sign of fusion but would have constituted a serious health hazard? Nuclear reactions generate nuclear radiation. That is the nature of the beast. Even without calculating exactly how much neutron radiation should have been emitted by four watts of fusion power, one indication that Pons and Fleischmann had observed too little was their reasonably robust appearance.* The radiation emitted from this level of power generation should have been sufficiently malign for its effects to have been noticeable. In fact, the radiation would have killed the two chem-

ists, not to mention seriously impaired the health of the students working nearby. But it hadn’t. So where were the neutrons? Bard and Schmidt ran a quick and dirty experiment that night. They

118 BH GARY

TAUBES

placed a palladium rod and a platinum rod ina flask of heavy water, hooked the two rods to an electric current, and compared the results of

that setup with those of a similar flask in which both electrodes were made of platinum. The idea was that if the deuterium fused in the palladium electrode, that cell would produce more heat than the cell that had two platinum electrodes. What Bard didn’t know was what manner of electrolyte Pons and Fleischmann had used. He tried D,SO,, a heavy

water variation of sulfuric acid. He and Schmidt ran both cells off the same power supply with about the same cell voltage, then looked for excessive heating in the palladium cell. “It would have to be something very big to see it,”’ Bard said. “We didn’t see anything.” Bard was marginally skeptical. He agreed with a philosophy that Fleischmann would later articulate: “that if you really don’t believe something deeply enough before you do an experiment, you will never get it to work.” It may sound like the kind of instructions the Good Witch gave Dorothy to get her home from Oz; nevertheless, one has to have some faith to get things going. By Saturday, Bard was calculating possible electrochemical routes to achieving fusion, searching for some understanding of what conditions might lead to very high pressure in the palladium. So he estimated how high the pressure would have to be to induce fusion, and came to 107° atmospheres, which is only a factor of 100 less than the 10?” atmospheres Fleischmann claimed they had achieved at Utah. This may have been an interesting coincidence, but it still seemed absurd. 10?” atmospheres 1s one of those enormous numbers that scientists bandy about with such a cavalier spirit: one billion billion billion times the pressure of the atmosphere on the surface of the earth. To put this enormous number in perspective, it is roughly 10 million billion times the pressure at the center of the sun. How could it be possible to produce even a few million atmospheres in any earthly apparatus without that apparatus being crushed or exploding? So Bard and Schmidt considered the possibility that microelectrodes— or rather ultramicroelectrodes—had to be used. Fleischmann and Pons had lately been working with ultramicroelectrodes, so this was not idle speculation. Bard thought that maybe with these “‘tiny, tiny electrodes,”’ they could induce a stupendous current density across an electrode and, in One instant, create an enormous pressure in the palladium. So they took twenty-five-micron wire and built ultramicroelectrodes of palladium. But they failed to detect any evidence of fusion.

BAD

SCIENCE

8 119

7 PASADENA

When Nate Lewis arrived at his Caltech lab on March 24 atalittle after 8:00 a.M., two of the postdocs—Mike Sailor and Reggie Penner— already had a cold fusion experiment on the verge of running. Lewis was skeptical. He told Penner and Sailor that the experiment wasn’t worth more than one day’s effort. It was also possible that Lewis, having a high opinion of his own talents, did not want to sweep up after Stan Pons and Martin Fleischmann, even if they had just discovered the greatest thing since fire. Lewis was proud of this reputation. ““You can go out and get a reading on me. I did things and got the respect of a fair number of people by going into an area that people have studied already, and doing things meticulously and seeing things that other people didn’t see. So I’m one of the people, where you say, ‘If Nate says it’s right, it’s right.’ ”” Lewis had done his doctorate at MIT under Mark Wrighton, who had been electrochemistry’s resident prodigy before Lewis came along. Then Lewis went to Stanford as an assistant professor without bothering with a postdoc. Now he was thirty-three, and tenured at Caltech. Lewis preferred a blend of scientific talents in his group. He believed that if he put an inorganic chemist in with an organic chemist in with a physical chemist and so on ideas from the different disciplines would cross-pollinate. His style was to publish eight papers at once on asingle subject, saturating it with these various disciplines. Thus, what Lewis published tended to be the last word. Chuck Martin said of Lewis: ‘“The thing you’ve got to realize about Nate is, underneath his calm exterior, he is one of the most fiercely competitive son-of-a-bitches you'll ever meet. He’s very focused, and he relentlessly pursues things.”’ Because of Lewis’s reputation and his science, he attracted postdoctoral and graduate students who wanted to pursue the most challenging science and thought they were good enough to handle it. In 1989 many of these happened to be Stanford émigrés who had come down with Lewis a year earlier, when Lewis jumped schools. Lewis gave them freedom to do what they wanted. Their long-term projects then evolved into publications. Lewis acted as mentor, impresario, and arbiter scientiarum. In the words of Mike Heben, one of the Stanford émigrés, Lewis

would “‘filter through the bullshit and find out what’s nght and what’s not. Mike Sailor had arrived in town with his Ph.D. from Northwestern 99>

120 © GARY

TAUBES

before gravitating to Lewis’s group. He was a synthetic chemist who had worked on metal cluster compounds. He said Lewis’s lab looked “like a fun place to spend a couple of years.” Reggie Penner had done his Ph.D. with Chuck Martin at Texas A&M. He was currently working with Mike Heben on a scanning tunneling microscope project. A commercial STM would have cost about $70,000, so Heben and Penner had built their own. “You can’t

buy a commercial instrument that will meet our requirements,’ Penner said. With this hand-built STM, they had achieved atomic resolution,

which is to say they could make out single atoms under the microscope. It was like a novice race car driver building his own Formula I because a Ferrari wouldn’t meet his needs. Penner and Sailor had both studied the lengthy coverage of cold fusion in the Los Angeles Times. Not being nuclear physicists, they didn’t know much about fusion, so they lacked a certain skepticism. In fact, they were, as Penner would say, blown away by the news. He was even alittle angry that he hadn’t thought of the idea himself. They were familiar with both Fleischmann and Pons, whom they knew to be pioneers in microelectrode research. Penner estimated that he had forty some papers in his files written by one or the other or the two together. He had met them both. When Lewis walked into his lab on the morning of the twenty-fourth, Sailor and Penner were assembling their first cold fusion cell. They had bootlegged a strip of palladium and purchased the heavy water from the stockroom—-sixty dollars for one hundred grams. They had assumed Lewis could afford it. Curiously, neither Sailor nor Penner expected to succeed. They assumed that the experiment had to be considerably more complicated than the Los Angeles Times had made it out to be. Sailor, however, wanted to tell his grandchildren that he had tried cold fusion the day after the discovery was announced. As the day went on, the Caltech experiment went through various stages. They ran a control cell by replacing heavy water with ordinary water, the idea being that if excess heat was produced by fusion of the deuterium molecules, then ordinary water, which is only 14000 deute-

rium, would certainly induce less fusion and less heat. This is the most fundamental commandment in the canon of experimental technique. To reach an unimpeachable conclusion establishing the cause of an effect, run controls. E. Bright Wilson, Jr., in his classic

1952 volume An Introduction to Scientific Research, described controls as “similar test specimens which are subjected to as nearly as possible the

BAD

SCIENCE

#8 121

same treatment as the objects of the experiment, except for the change in the variable under study.” The cell with ordinary water seemed to behave identically to the cell with heavy water. Neither of them seemed to have induced a noticeable nuclear fusion reaction. Because they expected neutrons and gamma rays to radiate vigorously from the cells if they induced fusion, they hung up radioactive warning signs and borrowed Geiger counters from a radiochemistry lab. They also placed the incipient fusion cells on top of Polaroid film, under the mistaken assumption that if gamma rays were emitted they would expose the film. Sailor said later, “Gamma rays would go right through the stuff, but we didn’t know that.’ Nothing happened. They joked about lining their shorts with lead foil. “Neutrons would go right through that stuff, too,”’ Sailor said. “But we didn’t know that either. Meanwhile, Lewis talked to John Gladysz, a chemist from the University of Utah who was on sabbatical for the semester at Caltech. Gladysz had known Pons for seven years and told Lewis that he’d ‘“‘bet on Stan.” Lewis also decided that Pons didn’t look like someone who “‘went off the deep end, . . . didn’t look like he was really incoherent.”’ Still, they might have quit right then had not a synergistic effect begun with the physics department. Early in the day, Steve Koonin called Lewis. At age thirty-seven, Koonin had as formidable a reputation in nuclear physics as Lewis had in electrochemistry. Koonin was on sabbatical from Caltech at the University of California at Santa Barbara and called to find out what Lewis knew of cold fusion and these two guys Pons and Fleischmann. “They’re a little flaky,” Lewis said, “but they’re not crazy. I don’t know what to think.” In return, Koonin told him that there already seemed to be an independent confirmation. He told him that Steve Jones at BYU had apparently detected neutrons emitted from the kind of electrolysis cells that Pons and Fleischmann had used. Koonin had chaired the JASON panel on Jones’s muon-catalyzed fusion work, although Jones had never mentioned cold fusion at the time. But Jones had been mentioned in the Los Angeles Times article that morning: Physicist Steven Jones of Brigham Young University in nearby Provo is

widely recognized for his work in a similar field which involves the use of subatomic particles called muons to create a fusion reaction. That generally is regarded as a promising area of research, but Jones indicated in a tele-

122 © GARY

TAUBES

phone interview Thursday that he has switched his research to the same area now being studied by Pons and Fleischmann. Jones, who was reluctant to even discuss his work until a formal paper is published in May, said that after several years of frustration, ‘‘we’re seeing something significant.’’

It certainly appeared asifJones, a recognized name in fusion physics, had seen Pons and Fleischmann’s work, found it compelling, and replicated it. The Times also quoted Jones saying, “‘It’s a scientific success,” which seemed definitive. Koonin suggested that Lewis connect with Charles Barnes and the physicists at the Kellogg Laboratory, the nuclear physics lab at Caltech. Lewis suddenly began looking at the cold fusion scenario as not quite so implausible. So he called Barnes, a sixty-seven-year-old astrophysicist whose forte was studying exotic fusion reactions in stars and who was on the verge of calling the head of the chemistry department to get a line on a good electrochemist. Barnes had already discussed cold fusion with two of his research fellows, Steve Kellogg and T. R. Wang. It seemed that Barnes and his colleagues had one of the best neutron detectors in the world, and if they could learn the proper electrochemistry, they could set up a cell and be in business. So Lewis made the arrangements with Barnes, then told Penner and Sailor that they had a neutron detector at their disposal. The two postdocs put their cells on a pushcart and wheeled them across the campus. Kellogg and Wang then inserted the cells into the center of the neutron detector, which they called the neutron polycube. It was made up of a dozen neutron detectors arranged in a twelve-inch-diameter cylindrical array. They all looked inward on a four-inch-square borehole. The entire contraption, complete with paraffin shielding to slow down any neutrons, was about the size of a phone booth or a shower stall. Kellogg and Wang placed the cells in the borehole and rolled the detector into its chamber. Then they waited. The cube was 100,000 times more sensitive than the neutron detector Pons and Fleischmann had used. So if these Caltech cells emitted any neutrons at all, the cube would have little trouble detecting them. It observed nothing. Meanwhile, Lewis spent much of the day calling around the electrochemistry community trying to learn the exact ingredients in Pons and Fleischmann’s electrochemical cells. He was primarily concerned with the nature of the electrolyte. Penner and Sailor were using perchlorate, and if that wasn’t part of the Utah recipe, he wanted it out. “It’s like

BAD

SCIENCE

8 123

rocket fuel,” Sailor observed. ‘“‘There are documented cases of perchlo-

rate exploding on people, people losing eyes and fingers. There’s a lot of chemical energy in perchlorate.” Lewis had been trying to reach Pons all day; by Friday evening he managed to get hold of Jim Brophy. Lewis told Brophy that they were trying to replicate the work; they needed help, and they would love to confirm this. He said he knew Stan Pons personally, and he added that they had one hell of a sensitive neutron detector, and they were already going full blast. Brophy said he’d see what he could do. Lewis even sent Pons a message through the electronic mail network, or e-mail for short: Date: Fri, 24 Mar 89 18:25:16

From: pgs%Xray. Caltech. Subject: FUSION To: pons Hi stan.

| am sure that you have been inundated with requests, but we have an extremely sensitive neutron counter and can confirm the reaction, however our initial attempts . . . did not yield any neutrons. Could you please either send me a preprint and/or bitnet explicit directions of electrode preparation, D electrolytes, voltages, etc., so that we can confirm the experiments and quantitate the neutron yield? The particle physicists here have been very helpful and we can certainly lend credence to the proposal if you would provide us with explicit directions. Thanks for your help. . . . Pons called Lewis back that night around 8:30. The Utah chemist sounded relaxed on the phone, surprisingly affable. The conversation, however, was a peculiar one. “Hey, Id like for you to confirm this,’ Pons said. “We'd love to do it,” said Lewis.

“ll send you a preprint,” Pons said. “But I can’t tell you about it now.” “Why not?” “I can’t,” Pons said. Then he added, ‘““We don’t think there’s many

neutrons. This is a nonclassical nuclear reaction.”’ “Oh.”’

“Don’t worry,” Pons said. “You'll figure it out. Look for the heat. The heat’s the ticket. And I'll send you a preprint when we can... . Be careful, we’ve had explosions. I don’t want you to hurt yourself. I want you to be careful.”

124 B GARY

TAUBES

‘Well, thanks a lot,”’ Lewis said, trying to sound sincere. “Well, what

can we do wrong? What should we avoid? Because I don’t want to hurt myself.” ‘Avoid high currents and sharp edges,” said Pons. “Okay,” Lewis said. That was it. By the end of the day, Lewis’s chemists and Barnes’s physicists agreed that they had barely enough information to make it worthwhile continuing. If any neutron radiation had been emitted by their cold fusion cells, they’d have seen it. However, if they had been on the verge of quitting earlier, by Friday night it appeared that cold fusion had its hooks in these Caltech scientists. Now, heaven help them, it seemed to be a matter of

pride. Saturday morning Sailor started building an apparatus called a calorimeter. Pons had said look for the excess heat. The calorimeter would measure the heat output from the cells.

8 CAMBRIDGE

The first cold fusion attempt at the Massachusetts Institute of Technology came from a disenfranchised fraternity turned cooperative living group. On Friday morning they circulated copies of an announcement that included a photograph ofa glass jar that looked as if it might once have held strawberry preserves. Now all one could see was a metal rod, or maybe two, emerging from the top. This jar sat in a tub of water, which sat within a pile of bricks. Beneath the photo, it said: Here’s a photo of MIT’s first attempt at a room temp fusion reactor, manufactured in 15 minutes last night. We did heavy water (99.8%) electrolysis with two palladium electrodes, temp in vial went from 20 degrees centigrade to 50 degrees centigrade. Doesn’t mean shit since the electrolysis process itself is exothermic. Sorry about the overexposed photo taken with an oscilloscope camera. That’s all we had available. (The lead bricks were for neutron shielding.) Just some undergraduates having fines

It was signed by a member of Pi Kappa Alpha, or pika, as the coop now referred to itself.

BAD

SCIENCE

8 125

Marcel Gaudreau, a nuclear engineer at the MIT Plasma Fusion Center, was handed a copy of the pika announcement on Friday morning. When he read it, he started to take this cold fusion business seriously. Gaudreau stopped in to see his boss, Ron Parker, who was the director

of the Plasma Fusion Center. Parker had already heard about cold fusion. He thought Pons and Fleischmann were what he called “squirrels from his nut file.”

The Plasma Fusion Center, known as the PFC, is one of the half-

dozen research centers in America dedicated to research in conventional thermonuclear fusion. It is located behind the MIT campus, in a part of Cambridge made up of reconstituted candy factories. Parker oversaw a $25-muillion-a-year program aimed at achieving asustained fusion reaction in a tokamak, an immense donut-shaped fusion reactor.® He expected that MIT’s next generation of tokamak would reach break even, which is the point at which the device would generate more energy than it took to keep it running. Parker was a serious man, despite a distinct resemblance to the late British comedian Peter Sellers. He had invested his life in fusion power and believed that the only things needed to turn it into a viable energy source were hard work and dedication. He hoped he would see success in his lifetime. Gaudreau was only slightly less serious, a frenetic French Canadian who had also devoted his life to fusion. At the time he met with Parker, Gaudreau’s working hypothesis was that Pons and Fleischmann’s little electrolytic cell was just a display model they could show the press. Gaudreau considered this a variation on what he called the Christmas tree effect: ‘Say for example you want to bring the president to the control room, turn all the terminals on. That’s called the Christmas tree effect, and the more light the better.”’ So this cell existed only to demonstrate the simplicity of the device. Meanwhile, somewhere in their basement laboratory, Pons and Fleischmann

had sequestered a huge reactor surrounded by tons of concrete shielding. If this were true, it would explain how cold fusion might have generated the requisite neutron radiation without having proven detrimental to the health of the two chemists. Gaudreau told Parker, “‘Ron, I’ve got enough things in my life nght now. I don’t need anything more. But if you need someone to look into cold fusion, I’m ready.”’ Gaudreau spent Friday night viewing videotapes of the news reports. He also called the pikas. He offered them “‘legitimacy”’ for their experiments. He’d give them a laboratory, neutron detectors, and safety equipment. In return, he insisted that they refrain from running any more cold

126

HBGARY

TAUBES

fusion experiments until he could check them out. Gaudreau wanted assurance that the pikas weren’t frying themselves and maybe even their neighbors with neutron radiation. The pikas agreed. First thing Saturday morning, Gaudreau stopped by the former fraternity house. He picked up the cold fusion jar and put it in a lead-lined bucket to carry it back to the PFC. We don’t know what this thing is doing, Gaudreau thought.

While Gaudreau gingerly walked across Cambridge with something that for all he knew might be a small untactical nuclear weapon, Mark Wrighton was meeting with a handful of his electrochemists and grudgingly deciding that they should pursue cold fusion. Wrighton was another legendary figure in electrochemistry. His career had been almost preternaturally accelerated. He was the son of a navy man and had moved around alot as a youth. He was born in Jacksonville, Florida, and was schooled in Virginia, Tennessee, Maryland, and Newfoundland. He had studied chemistry at Florida State,

graduated in December 1969, took up at Caltech in January 1970, and two years later, at age twenty-three, he had his doctorate and was already an assistant professor at MIT. By age twenty-eight he was the youngest tenured professor in the history of MIT, and ten years later he was head of the chemistry department. In the meantime, he had pioneered various approaches for converting sunlight to electricity or chemical energy. His peers said it certainly wouldn’t surprise them if Wrighton came home with a Nobel Prize someday. In public, Wnghton appeared rational, deliberate, and sometimes cold. And, like Nate Lewis, who had studied

under Wrighton, when Wrighton published, his papers were expected to be definitive. Wnrighton’s researchers were simply confused by the Utah announcement. They knew that the palladium-hydrogen system, as it was called, had been studied for years by electrochemists, who had worked on it for methods of storing hydrogen, for cars that run on hydrogen. Physicists had studied it for separating tritium and deuterium from nuclear reactors. The system was so well known that it was hard to imagine any new discovery coming out of it, let alone one of such magnitude. While they were discussing it, Martin Schloh, a Belgian graduate student, walked in and said he’d already done the experiment, which seemed to make up their minds to look into it. Schloh was on the verge of finishing his doctoral work on microelectrodes, which meant he was familiar with the work of Pons and Fleischmann. His father had flown in from Europe on Friday and over breakfast shown Schloh a copy of the

BAD

SCIENCE

#127

Financial Times article, which had more of the experimental details than any of the American publications. Schloh then went back to his lab and began setting up the experiment. Friday night he wrote in his logbook, “I’m starting the experiment now,’ in case the experiment blew up and took him with it,although he didn’t really expect that to happen. His setup was crude. He was looking to induce a meltdown. He put power to his cell, stood back, and nothing happened. He tried it again with a Geiger counter hoping to detect neutrons. He spent much of Friday night and into the morning sitting by his cell, waiting patiently for the Geiger counter to announce that neutrons had been emitted. With this news, Wrighton called Parker, who said that they were

about to give cold fusion a try at the PFC and would greatly appreciate the assistance of experienced electrochemists. So Schloh, Vince Cammarata, and Dave Albagli, both graduate students, and Dick Crooks, a

postdoc, loaded up a car with power supplies, electrodes, recorders, various bottles of electrolytes, and odds and ends of electrochemistry supplies and drove the half mile to the PFC. “We wanted to be the ones to prove it right,” said Cammarata later. ‘We went in wanting to believe it, but being skeptical enough to say that if we see something we're not going to believe it immediately. Although the physicists were saying there weren’t enough neutrons, well, that was a convincing argument. But we knew that Pons and Fleischmann are not physicists, so we knew that any measurement that they made, especially a nuclear measurement, is not going to be the best measurement possible. So maybe they could miss some neutrons. They said this could be some hitherto unknown nuclear process. Who knows? If it is an unknown process, maybe it doesn’t produce neutrons. You can always rationalize anything, and you could rationalize how there would be nuclear fusion, and why [Pons and Fleischmann] wouldn’t die. There are no precedents for it. One way or the other there has to be a definitive proof, and we wanted to be the ones to definitively prove it.” By Saturday afternoon as many as twenty physicists and chemists were working away on a single fusion cell at the PFC. Another dozen people had drifted by to watch. Gaudreau set up a neutron detector and a few Geiger counters to monitor radiation. They asked everyone to move away from the cell and hit the power. Nothing happened. By evening MIT’s radiation safety officers had dropped by and roped off a safety zone around the cell. They gave instructions: nobody goes within the rope while the cells are running. Gaudreau and company now set up the cell within an altar of lead bricks, beneath which they put the

128 © GARY

TAUBES

neutron counter. They had a meter connected to the counter, which

would move if neutrons were detected. Then an undergraduate with binoculars read the meter from behind the safety rope. His instructions were to pull the plug immediately if he observed the cell going critical. The cell didn’t go critical, but that didn’t overly discourage them. After all, why should they succeed? They had no idea what the current should be, the voltage, the electrolyte, anything. Both Wnghton and Parker had tried to call Stan Pons to get answers, but they hadn’t gotten through. Parker and Gaudreau speculated that they needed to create a sheet of plasma between the electrolyte and the palladium. ““We thought,” Gaudreau said, ‘‘what if we were to electrolyze until the palladium was fully saturated, and then put a huge pulse on it and then drive it in? Go in there and hit it with five or ten thousand volts. So that’s what we did.” By early Sunday morning, Gaudreau, Cammarata, and a few pikas were the only ones left. Gaudreau had acquired a high-voltage capacitor, and for five thousandth of a second he hit one of the cells with 7000 volts. For a brief moment, they must have felt like Dr. Frankenstein cranking the juice into his monster. It’s alive! It’s alive! The cell glowed with an eerie blue light, which signified that they were harmlessly ionizing the electrolyte. No neutrons emerged, however. At six in the morning, Easter Sunday, they gave up. Gaudreau and Cammarata agreed to catch a few hours of sleep, meet at 2:00 P.M., and try again. Cammarata made it back to the PFC promptly, but Gaudreau stood him up. At some point Saturday night Gaudreau decided that there was only one certain way to determine exactly what Pons and Fleischmann had done. “So basically on Sunday morning,” he explained in his staccato style, “I went home, slept for a few hours, until eleven, got up, picked up the phone, called the airport, found out when the next plane to Utah was, got my American Express and my suitcase. Took off.’’ Cammarata only found out on Monday when Gaudreau called him from a pay phone outside Jim Brophy’s office and left him a phone number. When Cammarata asked where Gaudreau had called from, Dave Albagli, who took the call, responded “wherever he is, the area code is 801.” Utah, of course.

Although Gaudreau was thirty-six when cold fusion entered his life, his routine energy level and his overall attitude had been compared, not unjustly, with those of a hyperactive child. He had come to MIT in 1972 from a small town outside Quebec City. His father was a biology profes-

BAD

SCIENCE

8 129

sor in a small college, and, at the time, Gaudreau spoke only French Canadian. “I can honestly say that I was just off the boat.’’ Now he wore an MIT

ring, on the inside of which, he was proud to admit, were

inscribed the Greek letters Theta Delta Chi. This was the symbol of both his Boy Scout troop and his fraternity, of which he was still an active participant. Ron Parker described Gaudreau as an interesting guy to talk to, “‘although in a lot of ways he’s not representative of MIT.” With the underclassmen whom Gaudreau referred to fondly as his fraternity brothers, he was building a full-scale tokamak fusion reactor. They had no official funding. Theta Delta Chi was, in the lingo, bootlegging a fusion reactor, an endeavor that did seem somehow representative of the highly competitive and cerebral atmosphere of MIT. As far as cold fusion went, Gaudreau refused to have any preconceived opinion. ‘Forget about believing,”’ he said. ‘Leave that to the church. Look at the facts. Figure it out.”’ If Gaudreau could understand hot fusion and build tokamaks, he was damn well going to figure out cold fusion: “Because ifI can’t figure out these little jobs, I’ve got no business building machines. If] can’t figure this thing out, you know, I should go into the garment business.”’

9 SALTLAKE

CITY.

GROUND

ZERO

Friday, March 24, Martin Fleischmann left for England, and Stan Pons

spent the weekend without his mentor in what constituted a state of electronic siege. Pons had told reporters at the press conference that he planned to go skiing, but it’s conceivable that he never got off the telephone. Many of these calls were from well-wishers, including Edward Teller,

the director emeritus of the Lawrence Livermore National Laboratory. Teller was the mind behind the hydrogen bomb and the Strategic Defense Initiative, which made him a controversial figure, but he was also a great believer in technology and was always looking for the unexpected.’ Teller thought cold fusion sounded promising, and on Friday he convened a task force at Livermore to delve into it. This news must have bolstered Pons’s faith that they had done the night thing. Pons also received a call from Carlo Rubbia, who had shared the 1984 Nobel Prize

130 © GARY

TAUBES

in physics and was director general of CERN, the huge European high-energy physics laboratory. This was a conference call arranged by the Italian newspaper Repubblica between Rubbia in Geneva, Fleischmann,

who was stranded momentarily in San Francisco, Pons in Salt

Lake City, and Repubblica reporters in Rome and New York. Rubbia took the opportunity to invite Fleischmann to speak at CERN. Once again Pons saw the call as confirmation. He said that the conversation was fantastic: ‘“The man is brilliant. He picked up immediately on some of the subtleties we were discussing.” On Monday morning, it seemed that every young electrochemist in the world wanted to work alongside the coinventor of cold fusion; every newspaper reporter wanted to interview him; every scientist wanted to know how he’d done it. The deluge forced Pons to change his home phone number almost immediately, and within days change it yet again. And he got several new fax numbers and had an unlisted number put in his laboratory so he could reach his own researchers. With all the press and all the calls, no one ever stopped to wonder how this would affect a man who admitted to stage fright while teaching his undergraduate chemistry classes. Throughout his life, Pons had displayed a marked aversion to strangers, faceless crowds, and bureaucracies. It may

have been a product of his childhood; his Waldensian ancestors had suffered through eight centuries of religious persecution, so feelings of persecution may have been endemic to their psyche. What was certain was that if Pons had a choice, he always preferred working with family and friends. He’d had his son, Joey, doing research for him, and Joey’s name appeared as coauthor on several of his papers. His wife, Sheila, did much of his secretarial work. (This was to cause some embarrassment later, when it was discovered that both his son and his wife were on his chemistry department payroll, although the work they were doing seemed legitimate.) His personal lawyer, C. Gary Triggs, was one of his oldest friends. Even when it came to funding for his research—Pons relied on the Office of Naval Research for $300,000 a year—his contacts

were Pete Schmidt, whom he’d known for twenty years, since they were undergraduates at Wake Forest University, and Bob Nowak, who had studied with Schmidt and had received his doctorate at Michigan under the same man, Harry Mark, who had been Pons’s adviser and mentor.® With cold fusion, Stan Pons became a public man, whether he liked

it or not. He seemed genuinely upset about the pledge of $5 million from the state of Utah, and he kept telling friends and administrators that all he wanted was to return to his laboratory and get back to work. He told The Deseret News that the single word he would use to describe himself

BAD

SCIENCE

# 131

was “‘scared.”’ The News, in turn, referred to him as ‘“‘the unpresuming U. chemistry professor, whose name may someday rank along side Einstein, Edison and Newton.”’

Meanwhile, events at the laboratory were becoming progressively weirder. After the press conference, Pons and Fleischmann had lost their transparencies, which they assumed had been stolen. Sunday night Pons apparently realized for the first time that his cold fusion lab books had also vanished. These happened to contain no more nor less than all the priceless data on the experiments. Whether Pons believed at this point that his graduate student Marvin Hawkins had stolen them is unclear, but Pons called a young postdoc named Mark Anderson and asked if he would take over the lab work on the cold fusion research. (Pons did not tell Anderson about the lab books, only that he had decided it was time Marvin Hawkins went back to his thesis work.) Hawkins at the time was in Park City, forty-five minutes up in the mountains, where he was skiing with his wife and children. As Hawkins tells it, he had left that morning, after Pons had graciously offered to cover his expenses. Then on Monday morning, Pons called him in Park City, telling him that he needed the lab books, and Hawkins said he

couldn’t get them. He said he had put them for safekeeping in his brother’s safe deposit box, and he wouldn’t be able to reach his brother before the end of banking hours. Hawkins was under the impression that after the disappearance of the transparencies, Pons and Fleischmann had discussed with him how to safeguard the original lab books, so he had made copies to keep in the lab and deposited the originals in the safe deposit box. After talking to Pons Monday morning, Hawkins said, he drove to Salt Lake City and left Pons a photocopy of the books and a note: “I couldn’t get to the safety deposit box today. I’ll call you later. Here is a complete copy of the books.” That seemed simple enough. Then, into this escalating weirdness appeared Marcel Gaudreau, fresh off the plane from MIT, looking for a collaboration and willing to settle for anything that would give him a handle on room-temperature fusion. Monday morning, March 27, Gaudreau walked over to the univer-

sity’s public relations office and from there was directed to Jim Brophy, the vice president for research, who was now the front man for the cold fusion research. Gaudreau’s first question to Brophy was whether what

they had seen on television was the Christmas tree effect or the real thing. And Brophy said it was real. Right there, Gaudreau was thinking, they had a ‘“‘quantum leap in understanding.” At this point, Brophy and Gaudreau engaged in a guarded exchange

132

PBGARY

TAUBES

of information, a pas de deux that seems as though it might have been choreographed by Abbott and Costello or, at least, Alphonse and Gaston. Gaudreau wanted to move on to technical questions, but neither he nor Brophy was an electrochemist, so he didn’t know what to ask, and Brophy wouldn’t have been able to answer him anyway. Brophy did have a copy of the paper, which he showed Gaudreau but wouldn’t let him copy. Gaudreau looked at it, memorized a piece of information, and handed it back. Then he excused himself, stepped outside, and called Cammarata back at MIT. “Listen,” Gaudreau said to Cammarata,

“I’m not an electrochemist

but this is what I read: it says something about palladium rods, you know, the size of a pencil. Now, tell me the questions I need to ask.”’ “Okay,”? Cammarata said, ‘‘find out what’s in the solution. Is there

any type of electrolyte?”’ Then Gaudreau hustled nonchalantly back to Brophy, who once again handed him the paper. Then back to a phone. When Gaudreau first called Cammarata, his MIT office was crowded with chemists. So Gau-

dreau had him move to a “‘secure”” phone, where they would not be overheard. For some reason, Gaudreau would talk only to Cammarata,

maybe because he was the only one who had gone the distance Saturday night at the Plasma Fusion Center and thus had become an honorary frat brother. “It says 0.1 molar L-I-O-D,” Gaudreau told Cammarata in another call. “Good, good, Marcel,” said Cammarata. ““Okay. Now find out what

current density for what voltage they use.”’ Eventually Gaudreau trotted over to the chemistry building looking for Stan Pons himself. He waited in the hallway outside the basement lab until Pons came out and then walked across campus with the coinventor of cold fusion. Pons was off to teach a class, which was a tribute to his

effort to maintain some normality. Gaudreau said, ‘‘] remember asking Pons if I could watch his class. He said, ‘I’d rather not, and you won’t

be impressed anyway.’ ”’ Pons told him to come back in the afternoon, which Gaudreau did,

along with Brophy and two other men Gaudreau thought “looked like lawyers. Nice pants and suits.”” While they waited, Hugo Rossi, dean of the College of Science, showed up. Gaudreau somehow knew that Rossi had studied at MIT. The way Rossi remembered it, Gaudreau wanted to launch an MITUtah cold fusion collaboration. Rossi was anxious to get Pons to work with some established group, although not necessarily Gaudreau’s. Rossi

BAD

SCIENCE

# 133

said Pons was put off by Gaudreau’s personality and aggressive nature. “Marcel told me that what he wanted to do was pack up Stan, his computers, his cell, and take the whole thing back to MIT,” Rossi said.

“It was clear that none of that was going to happen. So then what he wanted to do was just come and work as one of the people in Stan’s lab. I didn’t think that was so outrageous. I tried to promote that idea to Stan and see how he felt, but he had no intention of allowing it.”’ Finally Pons appeared and talked to Brophy in private, leaving the two men in the nice suits standing outside with Gaudreau. Then Brophy reappeared and departed with the two suits, leaving Pons to tell Gaudreau that he couldn’t see him after all. Something had happened, Pons said. He didn’t say what. He did promise that he would send a copy of the paper to MIT at the end of the week. Pons also said that the cold fusion work was eighteen months premature, but Gaudreau didn’t ask him what he meant by that. “You know, Dr. Pons,” Gaudreau said, “I know you're being pressured to come up with these neutrons, but if you have fusion without neutrons, it’d be a hell of a lot better thing. These neutrons are going to cause you grief.” To which Pons replied, as best Gaudreau could remember, “If they want neutrons, I can make ’em neutrons.”’

To Stan Pons, Gaudreau must have looked like Houdini, appearing and disappearing outside his lab throughout the week. Tuesday he was there, although he did nothing more than exchange pleasantries with Pons. He spent the next two days in Oak Ridge, Tennessee, where he had a scheduled course on remote robot manipulation at the Oak Ridge National Laboratories. Friday and Saturday he was back outside Pons’s door. After each brief contact with Pons, Gaudreau would run to a pay phone, call Ron Parker, and fill him in. Then he would hustle back,

hoping to find Pons somehow more accessible than before. Gaudreau said he never associated with the many reporters loitering outside the lab. If he saw any cameras or trench coats, he’d take the elevator down afloor, then shoot through one of the long corridors and try to approach Pons’s lab from a different direction. Or he would wait until the corridor was empty. The U had posted two security guards by the elevators outside the lab, and if one or the other wasn’t around, Gaudreau would sit in their chair. That way if someone asked him what

he was doing, he could say that he was waiting for Dr. Pons, which was true. “I was always very, very careful,’’ Gaudreau said. “I mean, Ireally thought that I was MIT’s representative.” Gaudreau left Utah one week after he arrived, when one of the senior

134 0 GARY

TAUBES

members of the Utah chemistry department called Mark Wnghton, who called Parker and suggested they get Gaudreau out of there before he got in trouble. At one point the University of Utah police sent a message to Pons’s lab referring to Gaudreau as the guy “‘who Stan can’t get nid of” and suggested that the next time he appeared, someone sneak away and call security, which will send someone out to “‘deal with the nuisance.” It’s possible that by the time this message was sent, Gaudreau was already on the plane home.

10

THE

PRESS

On the morning of March 27, The Wall Street Journal reported Pons and Fleischmann’s rationalization for the disturbing lack of neutrons. As the Journal explained, ‘‘Fusion physicists are accustomed to thinking of fusion reactions occurring in fractions of a second at enormously high temperatures and densities of hydrogen atoms. The reactions created by the two chemists take place over hours inside a solid crystal.” Thus, Pons said, ““There’s no reason the reaction [in the palladium] has to be the same’”’ as that seen in the run-of-the-mill fusion reactions with which nuclear physicists are familiar. From this evolved the working hypothesis for cold fusion that something almost magical happened to the fusion process within the cold molecular lattice of the palladium. This theory seems to stand in contradiction to one of the basic tenets of science, which is that the laws of nature here are the same as the laws

of nature elsewhere. The laws of conservation of energy and momentum, for example, apply equally on the moons of Jupiter, in the core of a neutron star or a red giant or a supernova, or on the Lexington Avenue subway at rush hour. Science traditionally progresses by expanding the consequences of those laws known to be true in familiar realms to those realms into which no one has yet been able to look. If the laws of physics were different inside a piece of palladium under electrolysis, it would open a Pandora’s box, at least as far as the pursuit of knowledge went. What if they were only different in some pieces of palladium and not in others? Or what if every piece had a different variation on a theme? Or maybe the laws only differed in one piece of palladium? Mine, for instance. This sounds absurd, but consider that

BAD

SCIENCE

#8 135

representatives of Johnson Matthey, the firm that supplied Pons and Fleischmann with palladium, took to talking of one particular preparation of the metal as “fusion grade.” Indeed, it could begin to sound suspiciously like magic. But once this whimsical notion was accepted as plausible, that remarkable and unique properties existed within the crystalline lattice of palladium, anything was possible. It may have been for this reason alone that the scientific community was soon split into believers and disbelievers. A belief in cold fusion required an act of faith from the start. And faith, traditionally, has had no

place in science, where unbridled skepticism is considered a virtue. So some had faith and became believers, or ‘‘true believers,”’ as they were

sometimes called, an appellation that was not intended to be complimentary. To believe in fusion without neutrons was to believe in a benevolent universe, or a benevolent God. Neither, traditionally, had ever been so kind.

11

Thanks

to Marcel

Gaudreau,

CAMBRIDGE

the MIT

scientists knew

on Monday,

March 27, that the Pons-Fleischmann electrolyte was lithium deuteroxide. ‘“That was like getting the first commandment,”’ Gaudreau said later, “when you had no commandments at all.”’? And they knew that the palladium rods were the size of a pencil, which gave them an idea of the scale of the apparatus. And they knew that the rods were supplied by a company called Johnson Matthey, which gave them three commandments by Gaudreau’s count. Dave Albagli and Martin Schloh called a Johnson Matthey subsidiary in New Hampshire and were told that the firm had only two sixmillimeter palladium rods in stock. So Mark Wrighton’s chemists weren’t the only parties who had become interested in palladium since the March 23 announcement. “Hold them,” Schloh said. ‘“‘Package them. We will be there night away.”’ Albagli and Schloh must have had an overpowering sense of urgency, because they got a speeding ticket on the drive. Still, when they arrived they were informed that they could get only one rod; the second had already been purchased by a quicker customer.

136 # GARY

TAUBES

Meanwhile, Mark Wrighton met with Ron Parker and Paul Linsay,

a former particle physicist who had recently signed on with Parker’s fusion group. Vince Cammarata and Dick Crooks from his own group rounded out the assembly. Wrighton said he was willing to do whatever was necessary to pursue cold fusion. His personal opinion seemed to be one of scientific impartiality. Wrighton believed that if cold fusion were true, ‘‘this revolutionizes everything and we are going to have to take a serious look.’’ And if they couldn’t be the first to discover, they could at least be the first to confirm. Wrighton offered the continued expertise of his chemists in return for the know-how of Parker’s physicists. Parker took him up on it. Wrighton had copies of the morning’s Wall Street Journal article on cold fusion, which included a warning from Stan Pons that these fusion experiments could be extremely dangerous: He said that in an early stage of the experiments the apparatus suddenly heated up to an estimated 5,000 degrees, destroying a laboratory hood and burning a four-inch-deep hole in the concrete floor.

Even Wnighton seemed amazed by the violence of this episode. Linsay, the physicist, asked him if he knew of any chemical reaction that could do that. Wnghton said no. So the physicists, who certainly had their doubts about the plausibility of cold fusion in terms of nuclear physics, were temporarily mollified. If the chairman of the MIT chemistry department couldn’t imagine how this explosion could come about, then maybe cold fusion had more validity than they thought. Monday night the Parker-Wrighton collaboration set up five cold fusion cells at the Plasma Fusion Center, in a room that had been built

to house MIT’s next generation of tokamak, called ALCATOR-C MOD. This reactor was expected to generate kilowatts of power from deuterium-deuterium fusion. That was not much power by the standards of a modern-day electricity generating plant, but it did represent a very lethal dose of neutron radiation. Hence, the room that would contain ALCATOR-C

MOD

was a cube, thirty feet on a side, with five-foot-

thick concrete walls to absorb the neutron bombardment. Portholes in the bunker allowed cables to run from the tokamak up to a control room that looked down on the cube from an angle. The five cold fusion cells were put on a table in this bunker. Thermometers'® were placed in the cells. Neutron and gamma ray detectors surrounded the cells. The researchers then set up their makeshift control station on the safe side of the portholes. This setup was what the MIT

BAD

SCIENCE

8 137

people referred to as their stage one calorimetry, a crude method of measuring the heat released by the cells. It would register fusion reactions if they were as dramatic as Pons and Fleischmann had claimed. It would probably not be sensitive enough to detect reactions that were considerably less dramatic. The chemists calculated that it would take roughly thirty hours to saturate their palladium electrodes with deuterium. So they set the cells charging and took notes. None of them had ever done calorimetry before—not the electrochemists or the physicists. They assumed that if the cells began generating watts of heat, their various instruments would show it. Stan Luckhardt, one of Parker’s physicists, said, “We were just

going to sit back and watch the temperature rise.”’ That this image has its absurd aspects did not escape notice by the participants. Not the least of the absurdities was the sight of this multimilhon-dollar building, awaiting its $100 million hot fusion reactor. There, bubbling away onatable in this barren bunker, was perhaps a thousand dollars’ worth of chemistry equipment that had been advertised to do the very same thing. And sitting outside were a handful of chemists and physicists staring through these portholes, rubbing their eyes and waiting for salvation. They were skeptical but excited nonetheless. Like agnostics at Lourdes, they didn’t particularly believe in miracles but wouldn’t have minded seeing one if the opportunity arose. Then there was the group from the materials science (metallurgy) department at MIT, which also wanted to do cold fusion experiments and had apparently been pointed in the direction of the PFC by the radiation safety office. They came by with file folders tucked under their arms bursting with Wall Street Journal articles and faxes. It was obvious that they had all the same information that Wnghton and Parker’s people did, with the exception of what Gaudreau had gathered out in Utah that morning. “They were just dying to get their hands on what was in the electrolyte,” said Dave Albagli. Wrighton’s chemists and Parker’s physicists wouldn’t tell them. Eventually, these interlopers were absorbed into the collaboration. The materials scientists were led by Ron Ballinger, who hadajoint appointment in nuclear engineering and materials research. For the chemists and physicists, Ballinger’s point of view turned out to be worth hearing. Apparently the two groups got to talking about Pons’s mysterious explosion/meltdown. Ballinger said he wasn’t as impressed as the physicists and chemists were. So what if they had an explosion? he asked. Ballinger was a very talkative guy, a natural raconteur, and he had all kinds of stories about palladium. He told his audience that palladium 1s

138 8 GARY

TAUBES

used as a catalyst in wood stoves, for instance, or the catalytic converters in automobiles. It is a surface on which chemical reactions will cheerfully take place. He said when you throw a palladium catalyst into a wood stove, you get what are called exothermic recombination reactions, and they can raise the temperature 300 to 400 degrees centigrade in a minute. If palladium happens to be in a hydride state, which is to say the hydrogen has bonded to the palladium molecules, and the palladium starts to erode and the hydrogen leaks out, this hydrogen could recombine with the oxygen in the air. It would be no problem whatsoever to make one of these electrodes glow red hot and melt. And he said that when a palladium rod becomes saturated with hydrogen, it expands. The hydrogen begins to leak out and recombines with the oxygen. And it does so on the surface of the palladium. Heat is released. The hotter it gets, the faster the hydrogen comes out, the more heat is released. And it has nowhere to go. Bang! Or the level of the electrolyte in the cell can drop, exposing the rod to the air. Bang! Hydrogen-oxygen recombination happens to be an extremely powerful chemical reaction. Space shuttles, for example, are launched into orbit

when their rocket engines recombine hydrogen and oxygen. This reaction, Ballinger explained to his audience, has one of the highest energy densities of any possible chemical reaction, which is to say it packs a bigger wallop than anything else you could mix in a laboratory. This is a rocket fuel, he said. What you need for a rocket is a big bang for your buck. (In fact, the potency of the effect in palladium had been recognized for 160-odd years, and it had even been used productively in cigarette lighters.) So maybe the mysterious explosion that Pons had mentioned that morning in The Wall Street Journal was not so mysterious. While Ballinger talked, the MIT cold fusion cells continued to charge. And as the palladium absorbed the deuterium, it began to expand. The heavy water began to evaporate. The rods began to blacken, fatten, and crack. After a while the scientists pulled out one rod that had been only partway submerged. “‘It looked,” said Paul Linsay, “‘like one of these ten-thousand-year-old fertility balls.’’

BAD

BRITISH

BOFFINS

BID FOR NUCLEAR

BLAZE

SCIENCE

8 139

OF GLORY

March 25, headline liverpool Daily Post ATOM

TEST MAN

SEEKS

SUN

SECRET

March 28, headline Northampton Chronicle & Echo

The Harwell laboratory is situated about fifty miles west of London on what used to be a Royal Air Force base in the idyllic British countryside. From the outside it looks like a sprawling industrial complex or maybe an architecturally uninspired institution for higher learning. However, visitors must be cleared through a guardhouse at the front gate before entering the property. These security measures are necessary because the better part of Harwell’s resources are dedicated to nuclear research. In 1986 the British government finished the transformation of Harwell into a commercial laboratory, funded through its customers in both industry and business, with the honorable goal of making money. Cold fusion, if it worked, would do just that. Martin Fleischmann had spent the long Easter weekend at home with his artist-wife in Tisbury, then announced on aradio interview that he would be at Harwell Tuesday morning to give a seminar. This prompted a plague of reporters to descend on Harwell, only to find that it was closed for the weekend and anyone who was in working wasn’t talking. At one point an Italian crew threatened to land a helicopter in the research director’s backyard if he wouldn’t give them an exclusive interview. The research director, Ron Bullough, apparently convinced them that an aerial assault would not work to their favor. For the most part, the British press had been marginally less optimistic about cold fusion than their new-world counterparts. Although the reporters had minimal access to the scientists at Harwell, these already had ten days’ experience in cold fusion and could curb enthusiasm. “It is not fair to encourage ridiculous optimism,” Bullough told the reporters, and for the most part they did not. Still, that the Harwell staff knew about cold fusion and was taking it seriously was evidence that the work might be credible. The resident scientists at Harwell assigned to the cold fusion beat were as curious as the media about what Fleischmann would have to say on Tuesday. Maybe more so. They had spent the weekend running cells and losing sleep. They had observed no neutrons coming from the cells and

140 © GARY

TAUBES

had spent Good Friday constructing more. The administration gave them carte blanche to take what they needed from the Harwell storerooms. With police assistance, they went around the stockrooms with a trolley helping themselves to anything they thought they might need. By midnight on Good Friday, halfa dozen cells were up and running, with neutron detectors and gamma ray detectors, and all the electrochemistry under computer control. The bulk of the work had been done by four scientists: two physicists, Martin Sené and David Findlay, and two

electrochemists,

Dave

Williams

and Derek

Craston.

These four

insisted that they weren’t overly anxious about the radiation hazard— “Well,” said Sené, ‘“we knew Fleischmann was alive’ —but the director

of Harwell was, and it was his laboratory. As Williams put it, the director had become ‘‘amazingly twitchy” about radiation. So they covered the cold fusion cells in plastic boxes, a precautionary measure should radioactive liquid suddenly squirt from the cells. They set up trip systems that would immediately shut the cells down should neutron and gamma rays begin radiating outward. They installed closed circuit television cameras to watch the meters on the radiation detectors, and those pictures were

fed back to a nearby accelerator control room, which was manned twenty-four hours a day. The accelerator operators were then givenalist of phone numbers and instructions: if the meters go above acertain level, phone these numbers and yell for help. Dave Williams was a longtime friend and collaborator of Martin Fleischmann, and he believed that Fleischmann was close to a genius, “‘a

fantastically creative guy.” If he said something worked, it most likely did work, even if other experts said it didn’t. Williams labored away at cold fusion cells until two in the morning on Saturday the twenty-fifth. Reporters woke him at eight in the morning. As he remembered it, “Some guy from the Daily Mirror wanted me to say that everyone was going to have alittle [cold fusion] heater in the corner of their living room, and that this is going to provide all the energy needs for the nation.” When Fleischmann arrived at Harwell on the morning of the twentyeighth, the day before his sixty-second birthday, he was confronted with an impressive gathering of the more expert researchers at the lab. His audience also included distinguished scientists from the UK’s Science Research Council, Central Electricity Generating Board, Atomic Energy Authority, and the huge magnetic confinement fusion program known as JET, for Joint European Torus, located nearby at the Culham Laboratory. “I have never seen so many FRS’s [Fellows of the Royal Society] gathered in one room in my life,” said Findlay.

BAD

SCIENCE

8 141

Fleischmann began his seminar with his heat data, or the “heat thing”’ as Williams later called it. Fleischmann flasheda series of tables and charts that seemed to demonstrate that the amount of heat emitted by the fusion cells was greater than anything they could explain by a chemical reaction. Fleischmann’s learned audience seemed to find the figures for heat generation impressive. Still, Fleischmann was asked if he had done control experiments with light water, and he responded that they hadn’t had time because of the frantic rush to go public.'! Ron Bullough wondered how Pons and Fleischmann could have supposed they had made a discovery if they hadn’t done any controls. With that, Bullough said, ‘he went down about five notches in my view.” Fleischmann turned next to his radiation data, beginning with what is known in the business as a gamma ray spectrum. These gamma rays were indirect evidence for the existence of neutrons but the primary evidence for fusion, as far as physicists were concerned. If the cell was emitting neutrons, then the neutrons would interact with the water in the sur-

rounding bath, and that interaction would cause gamma rays to be emitted. So Fleischmann showed his gamma ray spectrum, and Peter Iredale, who was director of the Harwell Laboratory, looked at Fleisch-

mann and said, “It’s wrong.”’ Then the other physicists in the room echoed his comment. To them it didn’t look like a gamma ray spectrum, and they knew what a gamma ray spectrum should look like. How Fleischmann felt about the pithy and unequivocal Harwell criticism can only be guessed. He said that he would call Stan Pons back in Salt Lake City and ask him about the spectrum. Derek Craston, who was there, recalled that even with a weekend of rest, ‘‘Fleischmann looked

like a very tired man.” Still, the gamma ray spectrum was not Pons and Fleischmann’s only evidence of fusion. The two had plenty more, especially the heat that had been generated, and the Harwell scientists had been impressed with that. And Fleischmann was still anxious to have the Harwell scientists reproduce his results. After the seminar, he spent an hour with Williams and Craston, describing how to do the cold fusion experiment. He even sketched a detailed schematic diagram of a fusion cell, to make absolutely certain that they would replicate his experiment correctly. That afternoon Fleischmann spoke with reporters and was remarkably frank. Cold fusion, he said, represented either “‘a Nobel Prize or a lot of egg on my face.”’ Fleischmann seemed to be momentarily pessimistic. Cold fusion, he admitted, could be nothing more than a “horrible chain of misinterpretations and accidents.”

142

© GARY

TAUBES 13

SALTLAKE

CITY.

GROUND

ZERO

On March 28, Marvin Hawkins was told that the police were after him. The twenty-seven-year-old graduate student, who would refer to cold fusion as ‘“‘my baby”’ and the fusion cells as his “‘little puppies,”” had been having one shock after another since his adviser and his university had decided to take cold fusion public. As Hawkins told it, two days before the announcement he learned that his name would not be on the cold fusion papers, nor would he be permitted to participate in the press conference. After doing at least a good part of the lab work on cold fusion, Hawkins had reason to expect he would be listed as an author on the seminal paper, which would have meant a share of the Nobel Prize and any royalties. Now his contribution would not even be acknowledged. ‘Stan said that the university had decided that I had to play a lowerkey role in the announcement,’ Hawkins explained. ““The lower-key situation was that I wasn’t going to be involved in the press release and that they decided not to include my name on the paper. In that stage of the game you’re kind of going, ‘Oh well, okay.” And then I was walking across the campus and it just dawned on me like a brick that they had written me out. That, if it was the university behind it, the university simply said this man isn’t high enough profile. He’s not good enough to be here. And indeed that’s exactly what happened.” Hard as this had been for him to accept, Hawkins convinced himself

that he should simply shrug it off. “I said, ‘Okay, I can deal with this.’ ”’ Hawkins was a good soldier. He even accepted his exclusion from the press conference, because he did not want to “throw any muck” on such an important occasion. Then there was the matter of the missing lab books, which Hawkins

said he put in his brother’s safety deposit box. Pons’s version of the story, as he told it to Hugo Rossi among others, was that Hawkins decided to give the books to the Church of Jesus Christ of Latter-day Saints as the modern equivalent of Joseph Smith’s tablets, maybe because Hawkins was disgruntled by his exclusion from posterity. In some versions of the story, Pons said Hawkins tried to sell the books to the Mormon church for a million dollars. After a week of frustration trying to retrieve the books, Pons gota call from the LDS headquarters, saying the books were there and would he please come down and claim them. Rossi tended to

BAD

SCIENCE

1 143

believe this version, but then Rossi at this time tended to believe Pons.

The bulk of the evidence supports Hawkins’s version. In any case, Tuesday, March 28, Pons arrived at the laboratory to find

only the photocopies. Hawkins was still in the mountains skiing with his family. He said he called Pons first thing in the morning to check in. “I had called immediately around eight-thirty, and [Pons] said, ‘But look, I’ve got to have these notebooks, the original data, and I have to have

those for the patents,’ and he’s going on and on about this. And Isaid, ‘Good grief, I didn’t realize that it was that big ofa thing.’ I said, ‘I’ll come down right now, and I’ll drive to my brother’s, and I’ll find out.’ But I didn’t have my brother’s phone number where he worked. . . . So we come down from Park City, my wife and I, and we go into Stan’s office, and he’s just absolutely livid. He said that there’s a warrant out for my arrest and if I had any sense about me I would get those books just right now, and he went on and on. I’m going, ‘Good grief! Get real!’ My wife is just in tears. This guy’s been threatening me. He’s going to put me in jail!” Hawkins spent a few hours finding a lawyer, who told him Pons had no case and added, “Please, I beg of you. Please let them arrest you. We'll both be wealthy the rest of our lives.’’ The lawyer, according to Hawkins, told him to keep the lab books as long as Pons had been given the photocopies, but Hawkins observed that “‘that sounds like a lawyer.”’ His wife told him to “‘be careful, or you'll get burned,” but Hawkins insisted Pons had been good to him, and he returned the books. “Now we’re in fine shape,”’ Pons supposedly told Hawkins. Hawkins then went down to the lab to find that he had been locked out. Pons had replaced him on the cold fusion project with Mark Anderson. Then Pons had instructed Anderson to remove all the palladium from Hawkins’s desk and have the locks changed on the laboratory door to keep Hawkins out. Anderson did both.’? Hawkins returned, infuriated, to Park City. Anderson then took to calling Hawkins for advice. “So he starts making four or five calls to me every day,’ Hawkins explained. ‘‘ ‘How do you do this? How do you that?’ And I was hurt, but I figured it’s still a bunch of bull.”

144 © GARY

TAUBES 14

AUSTIN

After working for four days to duplicate Pons and Fleischmann’s experiment, Al Bard was getting leery. He had already decided'to put together a more professional experiment, so he and Norman Schmidt had built a suitable electrolysis cell. They still, however, had little idea of what conditions were necessary to induce fusion. They had electrolyzed away, calculating how much heat was coming out of the cell and how much was going in. The accuracy wasn’t that good. The experiment would only detect large amounts of excess heat. Still, the watts out were always

fewer than the watts in. By Tuesday, March 28, Bard was trying to find neutron detectors. He called his old friend Doug Bennion, the BYU electrochemist. Bennion told him that Steve Jones, the BYU physicist, had also done the cold fusion experiment and had detected neutrons emitted. So now Bard knew that someone else had confirmed Pons and Fleischmann, or so it

seemed. Bennion said that Jones was using lithium deuteroxide for the electrolyte, just as Pons and Fleischmann had. Then, surprisingly, Bard heard from Stan Pons himself. The American Chemical Society had asked Bard to give an introductory talk at a special cold fusion session the society was throwing at their spring meeting in Dallas on April 12. Bard said he’d do it, but he wouldn’t talk in a vacuum. He had to talk to Pons and find out “what the heck he’s really done.” The ACS reached Pons and told him to call Bard, which he did,

and he gave Bard the details. He told him everything. It may have helped that by now Pons had heard from Fleischmann about the Harwell critique and knew they needed support. Pons told Bard that they had used a one-millimeter palladium rod, ten centimeters long, that they had set this within a cage of glass rods and wound a platinum wire around that. “Spacing is not critical,’’ Pons said, “but use about half a centimeter. Run it at 200 to 300 milliamps for 7.7 hours to initiate fusion.”’ Pons also told Bard that Jones had measured neutrons, confirming their work. And he told Bard about the meltdown. “So be careful,” Pons said.

Bard and Schmidt went back to the laboratory and began the cold fusion experiments once again, this time using the recipe prescribed by Pons himself. The new experiments, however, failed to produce signs of

fusion.

BAD TS

LOS

ALAMOS,

NEW

SCIENCE

#145

MEXICO

Immediately after the Utah announcement, the Los Alamos National Laboratory spontaneously blossomed with unauthorized cold fusion research. It started with two experiments, but soon a dozen teams of scientists were trying to duplicate cold fusion. The lab had been founded by the U.S. government in 1943 in the canyons and mesas above Santa Fe as the center of the Manhattan Project atomic bomb research. Since then it had evolved into forty-three square miles of national laboratory with a budget pushing a billion dollars a year. Some $60 million of this was spent annually on two conventional fusion programs. By March 28, the Los Alamos administration had officially appointed Rulon Linford to head up the lab’s cold fusion effort. Linford was director of the magnetic fusion energy program and head of the controlled thermonuclear research division. Coincidentally, he had obtained his bachelor’s in physics in 1966 at the University of Utah, where his father had once been head of the physics department. He had gotten his doctorate at MIT. With Linford coordinating cold fusion at Los Alamos, the contacts between the laboratory and Utah began to grow like spiderwebs. On the twenty-eighth, Linford heard from the governor’s office in Utah, which wanted to know what Los Alamos thought of cold fusion. Then he got a call from Ryszard Gajewski at the Department of Energy. Gajewski said he thought that Jones was right, and he had decided to fund Pons and Fleischmann, who had submitted a proposal to his office. Gajewski encouraged Linford to make contact with the University of Utah and see what he could do to help straighten things out. Meanwhile, Pons had his own contacts in the physical chemistry division at DOE, and they had already put him in contact with Robert Sherman, a physical chemist in the tritium division. It may have been through Sherman that Pons gathered the mistaken impression that Los Alamos had successfully reproduced his experiment, which is what he told the local papers. (U. CHEMIST BELIEVES FUSION CONFIRMED AT LOS ALAMOS, reported the March 28 Salt Lake Tribune, adding that Los Alamos officials would not confirm the report.) Tuesday afternoon, Linford called Pons and miraculously got through. Pons said that the powers-that-be in Utah were arguing against a collaboration with Los Alamos. According to Linford, “He articulated to me that these people wanted to obtain as much control over this discovery

146 8 GARY

TAUBES

and as much economic benefit [as they could], and to keep it within the state. They didn’t want a federal lab to take it away from them and run with it. [Pons’s] argument for wanting this contact was if it really was nuclear fusion, there may have been defense- or weapons-related spinoffs, or developments that could come from this that would clearly have not been controllable by the state of Utah. The government would step in anyway. It was better to have a lab like Los Alamos, which has all the knowledge and expertise, become

involved and make that assessment

quickly.” Linford asked Pons what the Utah chemist would like from Los Alamos, and Pons told him they needed help to understand exactly what it was they had discovered. On Wednesday, the twenty-ninth, Linford received an even stranger call, this one from an assistant secretary for security affairs at the Department of Energy. It seems this DOE fellow was “deeply apprehensive” about the weapons-related capabilities of cold fusion and wanted an informed opinion from Los Alamos. In particular, he wanted to know if the federal government should step in and classify the research. It seems that if cold fusion a la Pons and Fleischmann would emit copious neutron radiation, it would not be, in the words of one Caltech

graduate student, the benign little gizmo that it appeared to be. It would be a terrifying device, an ideal way of making plutonium in one’s basement. All that would be necessary would be ordinary uranium, which could be bought without too much difficulty. The neutrons would gradually convert it to bomb-grade plutonium, so for a few thousand dollars’ worth of heavy water, palladium, and platinum, anyone with passable expertise would be a long way toward building a nuclear bomb. Fortunately, the press hadn’t immediately noticed this aspect of the device, so it hadn’t made the papers. But it was there all the same. Linford had thought about this too. He told the assistant secretary that it was a little late to shut the barn door, as the technology had already been made public. Linford suggested that if federal agents were to swoop down and declare all cold fusion research top secret while the technology was being hailed by the press as salvation, the act would have nasty and explosive repercussions of its own.

BAD 16

SALT

LAKE

CITY.

SCIENCE

§ 147

PHYSICISTS

For six days now, physicists at the University of Utah had been receiving calls from reporters and scientists who had made the natural but incorrect assumption that any discovery in nuclear fusion would have come from the physics department, or at least that these physicists would have been consulted along the way. All the physicists could say was no, they knew no more about cold fusion than anyone else did, which was nothing; try calling Stan Pons. Several of the physicists had tried to see Pons as soon as he returned to work on Monday but with no success. Finally, Jim Brophy arranged for a delegation of physicists to get access to Pons. At eight in the morning on March 29, Craig Taylor, the head of the physics department, Mike Salamon, Gene Loh, and Orest Symko met with Brophy, who allowed them to study a copy of the paper that Pons and Fleischmann had submitted to the Journal of Electroanalytical Chemistry. Then the delegation walked over to the chemistry building and met with Pons, whom they found, to their pleasant surprise, warm, sensitive,

and patient. ““We spent an hour with him,” Salamon said. “He seemed like a very reasonable guy. Levelheaded.”’ Pons must have been making an effort, however, because Salamon and his colleagues politely pointed out that his evidence for cold fusion seemed to have potentially fatal loopholes. First there was the gamma ray spectrum. Salamon told Pons the same thing the Harwell physicists had said to Fleischmann—“‘It’s wrong.” Pons was not surprised by the news. “He was ready for it when we told him,” Salamon said.

The entire discussion continued along these disheartening lines. Frankly, the physicists told Pons, as far as the evidence for nuclear by-products—for neutrons and gamma rays—was concerned, they felt there were absolutely no data.’

17

PROVO

After March 23, the various administrators and scientists at BYU who

had been involved with the cold fusion negotiations felt, as Lamond Tullis, associate academic vice president, put it, that they had been

148 B GARY

TAUBES

“had.” Tullis said they were overcome by “righteous indignation” and an “enormous amount of stunned incredulity.” Jae Ballif, the provost, would say he felt ‘“devastated that, for whatever reason, agreements between honorable people were not kept.’’ Only Steve Jones professed to be relieved: now he no longer had to wonder what Pons and Fleisch. mann were up to. He knew, and it was done. What did move Jones to righteous indignation was the continuation of the “rumors” that he had engaged in scientific larceny. Jones heard from a friend at Los Alamos that Bob Sherman, who had been in touch

with Pons, was saying that Pons was saying that Jones had stolen their idea. He may have felt some stunned incredulity as well. Jones said, “I called Bob, andI said, ‘What’s going on here? If you have any questions, let me know. I’ll be glad to answer them.’ I told him what we had done. And he said, ‘I’m sorry.’ And that was the end of it.” Jones no longer professed to have doubts about the morality of his actions. He had been victimized by the press conference, and that was that. Pons, Fleischmann,

and Peterson were, to use the lingo of law

enforcement, the perpetrators. And Jones’s colleagues at BYU believed they had no reason to doubt Jones, because they too had been victimized by the press conference. Although they professed to be absolutely flummoxed by the duplicity of Chase Peterson, a man of previously unimpeachable integrity, they still were not so perplexed that they would question the credibility of Steve Jones. They just could not figure out how to reconcile this paradox. It was one of life’s mysteries: “I suppose [Pons, Fleischmann, Peterson] feel justified in what they did,” said Tullis, “because they apparently acted on the assumption that we stole their technology, which was a flagrantly ill-informed position. It doesn’t justify a total collapse in integrity.” On the day before the announcement, Jones began compiling his defense against future recurrences of the accusations. This was the official BYU history of cold fusion. It was Ballif’s idea, and Jones provided the necessary facts. Ballif said he couldn’t verify all the information Jones provided, but “I have no reason to doubt him.” This faith was further encouraged by Ryszard Gajewski at the Department of Energy, who spoke to the BYU administration the day after the news conference. Gajewski said he and his colleagues were equally appalled by the turn of events. “His advice to us,”’ said John Lamb, BYU’s director of research

administration, ““was to try to stay above the noise and the confusion, to continue our scientific work, to use the normal scientific procedures.”’ This was their inclination in any case, but it was nice to hear it echoed by DOE.

BAD

SCIENCE

§ 149

Meanwhile, Paul Richards, who was head of the BYU public com-

munications office, had to sit back and watch the media buy without question what Pons, Fleischmann, and company were selling. “Every day [there was] some new announcement,” Richards said, ‘‘and the press was falling for it hook, line, and sinker, and not confirming anything.”’ When the BYU work was mentioned, it was invariably taken as confirmation of Pons and Fleischmann. This conception, oddly enough, was propagated further by Gajewski. Reporters who called DOE for an official comment on cold fusion were directed to Gajewski, who said that he believed that fusion had been observed ina solid and that this was a “major scientific discovery.” Gajewski, of course, was referring to Jones’s work, but he would add that Pons and Fleischmann’s heat output was inexplicable, and he had decided to support the Utah research to the tune of $322,000. This seemed like a wholehearted endorsement of the

Pons-Fleischmann product by any standards.'* Until March 29, all Jones would tell reporters was that his work was independent and did not confirm the Utah results “in any way.” Jones would insist that he wanted to do this right and wait for publication before talking about his data. (He would, however, admit that the discrepancy between his neutrons and Pons and Fleischmann’s heat was a factor of a trillion.) All Richards could do was read callers the abstract that Jones had written for the American Physical Society meeting, which did not make for a satisfying story. Richards spent his time begging Jones to go public while fending off journalists who “were sniffing out that there was something fishy going on.” Then The Wall Street Journal arrived on the morning of the twenty-ninth and, said Richards, “really called a spade a spade.” SECOND FUSION DISCOVERY COMES TO LIGHT, read the headline. FINDINGS AT BRIGHAM

YOUNG

ARE SHROUDED

IN SECRECY,

SEEM LESS CONTRO-

VERSIAL. The story read in part: The Brigham Young scientists are refusing to talk about their findings until their report is published in a scientific journal, thereby avoiding the anger among researchers that the Salt Lake City scientists stirred up by announcing their discovery at a news conference Thursday.

The Journal also remarked that Johann Rafelski was developing a theory to explain the BYU results, which he hoped to “‘finalize” soon and submit to a scientific journal. Thus, Rafelski’s theory was “‘bolstering the credibility of the Brigham Young experiment,” which surely made it

150 #8 GARY

TAUBES

one of the few times in Rafelski’s career that his theoretical speculations had been cited as adding credibility to a subject. Richards finally invited the local television reporters down to Provo on March 29, and the reporters from The Deseret News and The Salt Lake Tribune the following day. When the press arrived, they were given a demonstration of the experiment and, of course, shown the lab note-

books. They were also shown Jones and Rafelski’s two-year-old Scientific American article, which was fortuitously entitled “Cold Nuclear Fusion,”

seeming to support Jones’s thesis that he was doing serious electrochemically induced fusion work long before Pons and Fleischmann’s proposal came his way. That the article was on muon-catalyzed fusion seemed to be irrelevant. To the layman, cold nuclear fusion was cold nuclear fusion. Jones reiterated to the reporters that the BYU results were much less dramatic than the Utah results. Did they promise energy salvation? ‘‘Not by a long shot,”’ Jones said, rolling his eyes. He then said that the odds that he and his colleagues had made a mistake were only “‘about one chance

in two

which

million,’

made

the BYU

results seem

only a

modicum less incontestable than the Pons-Fleischmann variety.’

18

‘‘OBSERVATION FUSION BY

IN

OF

COLD

CONDENSED

S.E—E.

JONES,

NUCLEAR

MATTER,’’ ETAL.

On March 29, Steve Jones’s Nature paper seemed to appear everywhere at once, thanks to the magic of the facsimile machine. Apparently Jones gave copies of the paper to half a dozen scientists, who then began photocopying and faxing it to friends and colleagues. Whoever was on the receiving end of the fax would make photocopies to be distributed to everyone in the neighborhood, then fax the article on to some new locale. “Observation

of Cold Nuclear

Fusion in Condensed

Matter,”

as

Jones’s paper was called, represented a distinct positive upturn in the course of cold fusion. ““This is a real paper and not a press release—it shows data,” wrote one awed researcher, who typed the good parts onto a computer bulletin board for those interested parties who had not been blessed by a fax.

BAD

SCIENCE

#15]

But those who read the BYU paper carefully were struck by how small

the effect was—two neutrons per hour. In fact, if it was, as The Wall Street

Journal had put it, “‘likely to be more immediately acceptable to scientists” than the Utah results, it was only because the effect was so close to zero as to be believable. It’s questionable whether the scientific community would have found Jones’s paper of interest at all had it not appeared before Pons and Fleischmann’s and thus given the information-starved community something to chew on. Chemists who found the minuscule BYU effect almost irrelevant nonetheless found Jones’s reporting of his electrolyte worthy of comment. It quickly became known as “‘the Mother Earth soup,” which is to say 1t seemed to contain a dozen different metal salts, possibly more: 0.2 g amounts Ees@z

.

H,O,

each: FeSO, NaSO,

=

- 7H,O,

10H,O,

(Calaly

NiCl,

- 6H,O,

(PO,)2

Oe

PdCh, Ti0OSO,

CaCO,, a

H,SO,

- 8H,O, and a very small amount of AuCN.

When this recipe was followed to the letter, the various ingredients would cross-react, oxidize, and reduce one another and either plate onto

the electrodes, creating what one chemist referred to as a “‘a mossy mess,” or settle, sludgelike, to the bottom.

The BYU

research, said David

Findlay of Harwell, “was probably not the divine last word in nuclear physics experiments, but you still felt a nagging doubt at the back of your mind that maybe there was some process going on there.”

19

PASADENA

Nate Lewis and his colleagues received the Jones paper on March 29 and immediately set about trying to duplicate that work. The Caltech chemists, however, were utterly mystified by the Mother Earth soup. They would refer to the electrolyzed concoction that resulted as, variously, a “‘piss-green solution with gray slag at the bottom” or “brownish red shit with this much garbage on the bottom,” or simply “‘a total mess.” So Lewis called Jones’s lab and was told not to use all the metal salts in each cell but only one or a few at a time. John Gladysz, who was on sabbatical at Caltech, discussed this in one of his e-mail notes home to

Utah. (He called these notes “‘reports from the mole.”) He wrote:

152 © GARY

TAUBES

Interestingly, that Jones paper is in deep disrepute. It reads well, but Nate has talked to Jones on the phone . . . there is an experiment that reads as though they got fusion in an eight component electrolysis mixture. What is really the case: they didn’t enter the brew into their lab books, all they know about the most successful expts you see quoted in their ms is that they contain “two, perhaps three, out of a possible eight components." Upon receiving the paper, Lewis and company immediately ran one Mother Earth cell, which emitted no neutrons. Lewis then decided to

opt for a technique akin to overkill. He calculated that, given the efficiency of the Caltech detector and the production of Jones’s cells—two neutrons per hour from as many as four to eight cells run simultaneously—they would need roughly fifty similar cells, all running at once, to assure that they would detect a neutron signal over the background. “So I said, We’re going to build fifty cells,’ Lewis later explained, “and they said, Ahhhrrrghh.” These researchers promptly dubbed what they were building the VLFA, for Very Large Fusion Array, which was a graduate student’s idea ofa pun, no insult intended, on the Very Large Array, a network of radio telescopes in the Southwest. “We put together this thing with fifty test tubes,” said Mike Sailor, “‘got this large sheet of titanium, analyzed it quickly with a scanning tunneling microscope. It was pure titanium. Cut it into fifty electrodes, wired it up, and covered the wire with epoxy, so

the wires wouldn’t get corroded in solution. Then the problem was that we had no good counterelectrodes. We didn’t have fifty wires of platinum or gold. We didn’t have that kind of money. But Charlie Barnes had this personal, research stock of gold. He just grabbed this chunk of gold and said, Go for it.”’ They also debated sacrificing their dental fillings, not to mention Lewis’s wedding ring, to the cause of cold fusion. They agreed that the fillings could go, but Lewis refused to part with the ring. ‘““We had some priorities still left,”’ he said. In any case, Sailor, Reggie Penner, Mike Heben, and company found a metal roller and spent the night of the twenty-ninth rolling Barnes’s chunk of gold into thin plates and slicing them into fifty electrodes. These they put in fifty tiny cold fusion cells; then they wired the entire array for fusion. They took it over to the Kellogg Radiation Lab and put it inside the cube. Although Pons and Fleischmann had been less than willing to tell anyone exactly how long one had to wait for fusion, Jones had stated clearly that his neutrons appeared within the first thirty minutes of

BAD

SCIENCE

§ 153

charging. So Lewis and company ran the fifty cells for several hours and saw nothing.” Meanwhile,

Lewis did receive an e-mail message from Pons in re-

sponse to one of his many requests for information, although it was of dubious assistance. The note suggested that Lewis ‘“‘should look for the effect first and the cause second,” and that was about it. Lewis observed

that the message was only slightly less abstruse than the choice of ingredients in the Mother Earth soup.

20

SALT

LAKE

CITY.

PHYSICISTS

AND

ADMINISTRATORS

At three o’clock in the afternoon on March 30, a delegation of Utah physicists took their critique of the cold fusion data to Chase Peterson. This time the delegation included Craig Taylor, the chairman of the department, Mike Salamon, Hugo Rossi, dean of the College of Science,

and Joe Taylor, academic vice president at the university. Taylor, who

was trained as a mathematician,

later took credit for

getting the physicists access to Peterson. He had become very pessimistic about cold fusion. He had found the press conference mildly embarrassing and, a few days later, was further dismayed when he finally learned exactly how big the purported discrepancy between heat output and neutrons was. “‘I didn’t understand that,” Taylor said, “until a few days

later when I was talking to Brophy. I asked him what exactly is the nature of the discrepancy, and he said it was a factor of about one billion. I nearly fell out of my chair. I remember telling myself, “We are in very deep shit.’ ”’ At the meeting with Peterson, Salamon did much of the talking. He told Peterson that as far as conventional nuclear physics was concerned, cold fusion was all but impossible and then explained that Pons and Fleischmann had a disconcerting dearth of evidence. Peterson asked whether it was conceivable that the experiment did create fusion but that the energy of the fusion was converted directly into heat and not into neutrons. That would explain why Pons and Fleischmann had observed so few, if any, of them. Craig Taylor (no relation to

154 § GARY

TAUBES

Joe Taylor) explained the rudimentary physics to Peterson: what we perceive of heat is the result of the vibration of entire molecules. A fusion reaction, by contrast, is what happens when the nuclei of two atoms fuse together. These two mechanisms exist on hugely different scales of distance and time. When fusion occurs, the newly formed nucleus will stabilize itself by ejecting a neutron or a proton in about 10 ~22, second (0.0000000000000000000001 seconds), which is blindingly fast. Atoms, by contrast, will vibrate about every 10 ~ 13 second. This is also quick by human standards, but it is one billion times slower than the decay of the nucleus, which is to say, the same ratio as thirty years to one second. As Taylor explained it to Peterson, the nucleus ejects this neutron with such instantaneous violence that everything else in the neighborhood appears to be virtually frozen in time. And the neutron, or whatever, is gone from the area well before the nearby molecules even know it existed. Expecting this particle’s energy to be somehow transmitted into heat, said Taylor, would be like shooting a pea at the speed of light out into our solar system and expecting it to noticeably warm up the planet Jupiter as it passes by. Under the circumstances, Salamon and Taylor argued, it was in the university’s best interest to take a very conservative and low-profile position on cold fusion. Peterson gave short shrift to the delegation’s advice. Perhaps it was too little advice too late, or a case of asking Peterson to close the barn doors now that he’d personally let the horses loose. Peterson had met the day before with Governor Bangerter, as well as the board of regents and various legislative leaders, in what the papers had called a “closed door session.”” Bangerter emerged announcing that the state legislature would hold a special cold fusion session on April 7, at which point he would urge them to appropriate the $5 million for cold fusion research. Peterson had announced that James C. Fletcher, who had just tendered his resignation as head of NASA, had decided to return triumphantly to the University of Utah, where he had once been president, to direct the cold fusion program. (Fletcher was interviewed in Washington by The Deseret News. With frightening syntax, he explained that cold fusion, if real, “would be the best thing that’s happened since I guess the atomic bomb—and that was bad, but in terms of breakthroughs it would certainly be a bigger factor than the transistor when that was built.’”” NASA then announced that Fletcher was not planning to run the Utah cold fusion effort and that Peterson, apparently, had been the victim of a misunderstanding.)

BAD

SCIENCE

#8 155

Peterson was not about to ask the governor to take back his $5 million offer. He did tell his physicists that he would couch the university’s hopes for cold fusion in a more cautious and conservative tone. And, for a day or so, he did.

“I admit we could be wrong,” Peterson told reporters. ‘“The science is good. It [the fusion experiment] is being repeated, we’re told, in labs in the U.S. and overseas. But there could be a hangup, a glitch, that we don’t know about.” To this statement, The Deseret News added: “It’s possible a reactor could prove unsafe, for reasons not now foreseen, or

that the energy produced could be so costly as to be economically

infeasible. If the commercial research comes to such a roadblock, work

will stop, Peterson promised, and no more money will be spent.” Jim Brophy also tried publicly to contain his enthusiasm after the physicists voiced their doubts. That same afternoon Brophy gave a memorable interview to The Deseret News: “If we had announced a breakthrough in energy-producing events,”’ he said, ‘no one would have shown up [at the press conference].”” He then added, “I don’t mean to say we used fusion falsely. There is fusion going on in the platinum lattice.” And then, while not admitting that they might be wrong, he did address the possible repercussions to the university’s esteem in the field if they were: “But there may be a taint that we can’t quite wash off. A graduate student may decide he’d rather go somewhere else or something, but there’s nothing we can’t handle.” What made this remark particularly poignant was that the rumor mills were already tying cold fusion to an earlier Utah scientific debacle. This was the X-ray laser affair of 1972. Although this had been a decade before the arrival of Stan Pons, the rumors spitefully connected him to that fiasco as well. There were, indeed, some disquieting similarities between the X-ray laser and cold fusion. The first was the brainchild of John Kepros, a postdoc at the time, working under chemistry professor Edward Eyring.'* Kepros and Eyring believed that by firing an ordinary infrared laser into a copper gel “sandwich” they could generate an X-ray laser out the other side. The press caught on to the discovery, and it made the papers, including such discerning journals as Newsweek. This prompted a circa 1972 delegation of Utah physicists to march on the administration and condemn the chemists for going public with a physics discovery that they seemed to understand not at all. In fact, no known physical effect could

explain how such a gel could produce a coherent beam of X-ray light. It was soon demonstrated, however, that the data could be explained

without invoking X-ray lasers.

156 @ GARY

TAUBES

The X-ray laser affair may have spawned the disparaging term Utah effect, which was thereafter applied freely to any public relations disaster originating within the state. (Cynics were already suggesting that cold fusion was the latest product of the Utah effect.) Still, the X-ray laser affair did not strike Peterson as a particularly portentous allegory. He later told the state legislature that after the X-ray laser fiasco the national ranking of Utah’s chemistry department rose dramatically. Hence, he said, ‘we should not operate from the premise that [early action] will embarrass us.”

21

THE

NETWORK

Without a doubt, the dominant factor in the first week of the flowering field of cold fusion research was the lack of information. Martin Fleischmann estimated that in that time he had given details of his experiment over the telephone to several hundred scientists. Yet the available information was depressingly unreliable. Larry Faulkner of the University of Illinois, for one, estimated that in the first few weeks after the Utah

announcement he spent some 30 percent of his time on the phone comparing cold fusion notes: ““The consistency of what we were getting out of Utah was extremely poor,” he said. After the telephone and fax machines, the major source of cold fusion information was the various computer networks—in particular, Usenet and Bitnet—which link universities, industry, and government facilities

throughout the world. These networks had begun carrying requests for cold fusion information within hours of the March 23 announcement. It was, as writer and physician Lewis Thomas would put it, a “‘collective derangement of minds in total disorder,” played out on neurons of semiconductors. The net spread a combination of information, rumors, requests, and reports. The first Financial Times article, for instance, had

been typed into the “‘net.”’ It had hard facts, provided by Fleischmann and missing from other newspaper reports. And, of course, because Stan Pons had a computer and was connected through the network to the outside world, his electronic mail address had been disseminated on the

net as well:

BAD

SCIENCE

8 157

Here's Dr. Pons’s e-address: pons(@ chemistry.utah.edu You can try to finger it first. Two

physicists, who requested to remain anonymous,

did try to

“finger’’ Pons’s mail account. To assure that they wouldn’t be traced,

they used the account of another physicist who rarely used his computer. They said later that they were “sweating like thieves breaking into a bank vault,” expecting the police. Pons’s account was guarded by a password of his choosing. The two physicists tried ‘‘cfusion,”” which didn’t work, then remembered that Pons had said he was a homebody, so they tried the names of several family members and hit eventually and correctly on Sheila, the name of his wife. So they broke into his computer account and spent two hours reading his mail. They were hoping for any clue: perhaps a letter to Fleischmann saying Pons had tried the palladiumtungsten mixture or some such and it worked, or maybe an order form to Johnson Matthey, the suppliers of the palladium. They justified the breaking and entering by the importance of the controversy and the inability to get a straight story out of Utah. One of the two said they found “‘junk, nothing of interest, crap.” He added that he knew of at least two other scientists who did the same, one from Oak

Ridge and one from Caltech. In any case, Pons later changed his password. Almost immediately after the Utah announcement, one could find on the network a variety of encyclopedic entries on deuterium and heavy water, as well as a lengthy compendium entitled “Everything you wanted to know about palladium and were afraid to ask,’ apparently cribbed from a book called Guide to Uncommon Metals by Eric N. Simons.’? The net also ran no shortage of theories to explain cold fusion should the experimental details pan out. This left an archaeological flowchart of the cold fusion speculation of at least one small segment of the scientific community. One entry, for instance, from Eugene Brooks, a physicist at the Lawrence Livermore Laboratory was entered at 00:45:36 GMT on March 26, which would make it midafternoon on Saturday in California. Brooks proposed a “solid-state process,” in which deuterium atoms are sucked into the lattice of the palladium and maybe fuse by quantum mechanical tunneling. Two hours later Paul Dietz, a computer scientist at the University of Rochester, signed on. He began, “Eugene Brooks proposed ina previous message a mechanism for how solid-state fusion might work,” and

158 @ GARY

TAUBES

explained his theory, which required that the palladium lattice be filled to the brim with deuterons. At that point, when one more deuteron entered, it would have to share a space somehow with another deuteron,

or else, Dietz speculated, “pry open” the palladium lattice, an act of violence that seemed “‘energetically expensive.”’ Thus, fusion might only be induced after all the holes in the palladium lattice are already filled. Following Dietz by some eighty minutes was Matt Kennel of Princeton, who began “‘[Paul Dietz] writes: Eugene Brooks proposed . . .” Kennel then described his own variation, ending with a modest “How does that sound?” Maybe it’s possible that electrochemistry people have been having fusion all along with Pd experiments (in regular water, the rate would be lower) but nobody would be crazy enough to look for gamma rays, for god’s sake! All three variations constituted what is disparagingly described as hand waving and required miracles of sorts to overcome the repulsion between the two positively charged deuterium nuclei, but such is the nature of speculation. Kennel was followed by a University of Texas astronomer (14:41:49 GMT) who suggested, incorrectly as it turned out, that perhaps the reason Pons and Fleischmann didn’t see neutrons was that their neutrons were absorbed in the palladium and heavy water. The Texan was then followed by a retransmission of a message from the twenty-fourth in which the correspondent reported, also incorrectly, that “the two guys who discovered the process are, together, the most respected electro-

chemists in the world.” And so it went, twenty-four hours a day, every day, for months.

22

SALTLAKE

GENEVA,

NEW

CITY,

YORK

CITY

March 31 promised a break in the general level of ignorance regarding who had done what with cold fusion, and how. Stan Pons was scheduled to give a cold fusion seminar at the U; Martin Fleischmann was booked

into the main auditorium at CERN, the huge European physics labora-

BAD

SCIENCE

8 159

tory outside Geneva; andSteve Jones was lecturing at Columbia University in New York. All three scientists played to packed houses. The Utah seminar was held in the auditorium of the chemistry building, which seats approximately 400. Pons had not wanted the affair to become any more of a circus than necessary, so the administration agreed, after some negotiation, not to hold it in the university’s basketball arena. It was estimated that a thousand people appeared hoping to hear Pons speak. Those who did get in had arrived at least forty-five minutes early. The local television crews were on hand, as well as NBC national news. Pons, however,

would not allow video recordings, and the reporters had to settle for interviewing members of the audience as they filed out.?° At CERN, which is the preeminent physics laboratory in the world, a similar madness ensued. There, however, those who got into Fleisch-

mann’s seminar only had to arrive thirty minutes early. Then Carlo Rubbia, the lab’s director general, refused to begin before ejecting the swarm of television and newspaper reporters. Rubbia explained that this was a scientific meeting, and he promised the reporters that a press conference would follow. As for Jones, the organizers of his Columbia seminar had to relocate it three times before finding an auditorium that could handle the crowd. Still, it was standing room only. Amiya Sen, the Columbia physicist who had invited Jones, described the event in familiar terms: ‘‘Of course, when he arrived on campus,” Sen said, “all hell broke loose. It was a

circus.” At Utah, Pons said that they had produced twenty-six watts of heat for every cubic centimeter of electrode and that one cell produced four megajoules of heat in 120 hours.’ All of this sounded awe-inspiring. He also admitted that they had seen a trillionfold too few neutrons to explain this mysterious heat and suggested therefore that the heat could not be coming entirely from deuterium-deuterium fusion, although no chemical phenomenon could explain this heat either. Pons said that the results were as puzzling to him as they were to everyone else and joked that they might see a viable commercial cold fusion technology in a hundred years. (In the early days of cold fusion, it was when Pons seemed to be joking that he seems to have been most honest.) One physicist standing in the rain after the seminar told reporters, ““Physicists say that the excess heat must be due to chemistry, the chemists maintain that it must be nuclear physics.” Rulon Linford of Los Alamos had been invited by Jim Brophy to attend the seminar. Linford found the lecture puzzling more than reveal-

160 8 GARY

TAUBES

ing. ‘‘It was short,” he said later, ‘‘and it was essentially philosophy: a table of results, of amounts, one picture of the flask in which the expenments were done, and some discussion of how the calorimeter was used.

But there was never any raw data presented during the lecture. You came away with some idea of what he was claiming, but no understanding of some of the care, or amount of data they had, or any of those kinds of things that allowed you to really judge the quality of the work.” Linford added, however,

“‘Not being an electrochemist myself, I don’t know

what’s usual in that field.” Fleischmann’s seminar at CERN was little more informative but still seemed eminently scientific to the physicists present, few of whom were experts in either solid state or nuclear physics. Rubbia, for instance, who had his own history of announcing discoveries without sufficient supporting evidence, was asked his opinion, and said, ‘“Dr. Fleischmann has

planted a seed—will the seed grow up? I think yes.” At Columbia, Jones seemed more interested in the past than in the future. He spent most of his time discussing his muon-catalyzed fusion results, which were the original subject of the seminar, then spent most of the remaining minutes on cold fusion, showing photocopies of his three-year-old notebooks as evidence that he had not pirated the idea from Pons and Fleischmann. He did emphasize that his neutron rate was the equivalent of one trillionth of Pons and Fleischmann’s heat emissions, and that Pons and Fleischmann’s evidence for neutrons was meaningless. His cold fusion results, he said, did not confirm Pons and Fleischmann’s research. He had observed neutrons from cold fusion; Pons and Fleisch-

mann had not. The seminar was entitled “First Detection of Cold Fusion Neutrons,” which said it all.

This became Jones’s stock lecture, which he would repeat with minor variations in Sicily, Geneva, Santa Barbara, Baltimore, and Los Angeles in that order. In Santa Barbara, for instance, Steve Koonin of Caltech

found his lecture “‘really disappointing. He spent the first half basically claim-staking and I found that really inappropriate.’’ One observer in Los Angeles said, “Jones, to everybody’s disgust, spent rather more time showing overheads of notarized pages in his lab book to establish prece~ dence then he did actually talking about what he found in his experiments. That went over like a lead balloon. Everybody was saying, ‘We don’t give a shit who thought about it first. We want to know whether it works or not.’ ’’ It was the same story, but Jones did care about not

only establishing precedence but distancing his results from those of Pons and Fleischmann. After all, he believed his were right. Richard Garwin of IBM was at the Columbia seminar. Garwin had

BAD

SCIENCE

#161

long before attained something akin to mythical standing in physics, as well as in the defense community. He had studied under Enrico Fermi at the University of Chicago. Fermi, who may have been the greatest physicist of the twentieth century, Einstein notwithstanding, called Garwin “the only true genius I’ve ever known.” For his postdoc, Garwin built the first hydrogen bomb, apparently over his 1951 summer vacation. Hans Bethe, the Tolstoy of quantum mechanics, said of Garwin that

he “is probably the smartest person I know,” then noted that the only other person he knew with Garwin’s level of talent was Fermi. William Happer of Princeton, who was director of JASON, called Garwin ‘“‘a perfect experimentalist.” He was said to have “exquisite common sense.” All of this made Garwin someone whose opinion one should weigh seriously, whether one agreed with it or not. Ryszard Gajewski had called Garwin just before the Utah announcement, informed him of the Utah-BYU contretemps, and asked him to be on a Department of Energy review panel that would assess the conflict. After March 23 that had become a moot issue, but Gajewski had sent Garwin the Jones preprint, and Garwin had decided to go to the seminar. As far as the evidence for cold fusion went, said Garwin succinctly,

“Jones didn’t have enough data.”’ Still, the audience in general, scientists and reporters alike, seemed to be less discriminating. ““There were a few worried faces in the audience at the start of the talk,’’ wrote one corre-

spondent on the network; “‘at the end, the speaker received a raucous applause.” The Washington Post reported that this was the first time “skeptical scientists and students” were presented with the results of a cold fusion experiment: “both appeared to be convinced that the fusion is real if not impressively large.”

23

‘‘ELECTROCHEMICALLY

NUCLEAR a]

MARTIN

FUSION

OF

FLEISCHMANN

INDUCED

DEUTERIUM, ’’ AND

STANLEY

PONS

It is possible to find the truth without controls, but the process has been demonstrated again and again to be notably

162 B GARY

TAUBES

inefficient, so that years may be required before it is appreciated that a given treatment is worthless. E. BRIGHT WILSON, An Introduction to Scientific Research

The Pons-Fleischmann paper in the Journal ofElectroanalytical Chemistry emerged from facsimile machines around the country just as the two chemists were giving their seminars. Ron Parker of MIT called it “a gift from God.” The paper, of course, was copied and sent through the system of faxes, distributing His bounty far and wide. The Wall Street Journal received a copy and announced the news in a headline the following Monday: HEAT SOURCE IN FUSION FIND MAY BE MYSTERY REACTION—FUSION PAPER SHEDS LIGHT ON FIND, EASING SOME SCIENTISTS’ SKEPTICISM. The Journal reported that Pons had provided copies of his paper to five scientists, and from those five it had propagated prodigiously.” In fact, the Utah paper, like the BYU paper two days before, was disappointing. It was, said Jim Brophy, “‘better than a telephone conversation, there’s no question about that,”’ which was certainly true. But that

was about it. Al Bard said the paper read like a “rush job,” and Dick Garwin called it “the worst so-called scientific paper” he had ever seen, adding that if the paper included the full extent of Pons and Fleischmann’s data, the two chemists needed a miracle.

Pons, Peterson, and Brophy would later talk up the paper as though it had been peer-reviewed, but that had not been the case. Roger Parsons, the managing editor of the Journal of Electroanalytical Chemistry, had served as the sole referee. He later admitted cheerfully that he had accepted it because Fleischmann was his longtime friend and he respected him asa scientist. And the paper undeniably was newsworthy. ““You take a chance,”’ Parsons observed. “I suppose that I could say that I took too much of a chance on this one.’’ The paper’s most glaring and immediate deficiency was that it made no mention of any control experiments. This raised the pithy question of how Pons and Fleischmann could have come to any conclusion about their experiments, let alone such a remarkable one as room-temperature fusion. As E. Bight Wilson phrased it in An Introduction to Scientific Research thirty-seven years before cold fusion: ‘If one doubts the necessity for controls, reflect on the statement: ‘It has been conclusively demonstrated by hundreds of experiments that the beating of tom-toms will restore the sun after an eclipse.’ ”’

BAD

SCIENCE

# 163

Laura Garwin, who is Dick Garwin’s daughter, was an editor at Nature

and would oversee Pons and Fleischmann’s submission to that journal. She said of the JEAC paper: “I was extremely surprised that that paper got published. It didn’t have the elementary control experiment. The obvious thing you see when you look at that paper is, Why didn’t they do it with H,O? Any high school student could’ve refereed it, because

of that obvious ingredient of the scientific method.’ Nate Lewis said, “You read the paper and you assume that there are controls done, and they did these things by rational means. Then you find out it wasn’t always that way.” The absence of controls would haunt the cold fusion episode like a recurring nightmare. Also conspicuously absent from the Pons-Fleischmann JEAC paper was raw data of any kind. All the numbers seemed to have passed through at least one level of interpretation and analysis. It was difficult to determine exactly what that analysis had been, and thus what the original data had been. On the one hand was the evidence for the by-products of a nuclear reactlon—tritium, neutrons, and gamma rays. Even if one believed what

Pons and Fleischmann reported in the paper, the numbers were so small as to be almost irrelevant.** “‘As soon as you read the paper,”’ remarked Larry Faulkner of the University of Illinois, “‘you realize the only thing they’d ever done was calorimetry.”’ Faulkner’s point was that even if they had the | tritium, neutron, and gamma ray work right, “‘the radiation was tremendously small compared to the heat effects.”’ Fleischmann himself was now telling inquirers that their observations stood or fell on the production of heat. He didn’t understand, he said, why everyone was off looking for the production of neutrons. He called it “a very barren search.” Fleischmann may have been right, but it was still virtually impossible to judge just how much heat Pons and Fleischmann’s cells had produced. The paper contained two tables that itemized the amount of heat produced by the cells. All the numbers were given as a rate of ‘‘excess heating.” In table 1, excess heating was set down as a function of the current density and electrode size. In table 2, which was the more awe-inspiring, excess heating was recorded as a percentage of break even, that being the point at which the energy going to charge and run the cell was equal to the energy, in the form of heat, produced by the cell. Of the three data columns of table 2, the third was the impressive one, although its meaning came through with less than crystal clarity. Here

164 © GARY

TAUBES

Table 2 Generation of excess enthalpy in Pd rod cathodes expressed as a percentage of break-even values. All percentages are based on 7D + 7D reactions, i.e., no projection to 7D + °T reactions Excess Heating/ Electrode Type

Rods

% of Break-Even

Dimensions /cm

Current Density /mA cm~?

a

0.1 x 10

8 64 512 8 64 S12 8 64 512

23 19 5 62 46 14 111 66 59

0.2 x 10

0.4 x 10

;

12 i 5 27 Zo zt 53 45 48

60 79 81 286 247 189 1224 438 839

*% of break-even based on Joule heat supplied to cell and anode reaction 4 OD” + 2D,0 + O, a 4te, ° % of break-even based on total energy supplied to cell and anode reaction 4 OD~ => 2 D,O + Os sane 3 “% of break-even based on total energy supplied to cell and for an electrode reaction Det ZODs > 2D,0 + 4e° with a cell potential of 0.5 V.

were the tremendous numbers for excess heating: 1224 percent of break even, 839 percent, 438 percent, 286 percent, and so on.

This was not raw data, which is an interesting term if one gives it a moment’s thought. Raw data has no particular positive or negative connotations. Yet it can be argued that if data is not ‘‘raw’’ then it is “cooked,” which does carry negative connotations. So be it. Scientists might have been better served had Pons and Fleischmann chosen to provide the raw data rather than this “interpreted” data. It would take some scientists weeks to figure out where these numbers came from; others, apparently, never would. One assumed that somehow, somewhere Pons and Fleischmann had measured these numbers. The paper didn’t seem to say.”° The Wrighton-Parker cold fusion collaborators at MIT read the paper simultaneously. As Dave Albagli told it, their unanimous and near immediate response to reading it was, Where’s the evidence? Why are we

BAD

SCIENCE

§ 165

doing this? “You hear they’ve been doing this for five years,” Albagli said, “‘and you say, Why don’t they already have it down? Why isn’t this just a brief Nobel Prize-winning paper that says it all, very succinctly and very emphatically?” There were three possible answers to these questions. The first and simplest was that Pons and Fleischmann had no evidence and cold fusion was a canard. The second was that a man of Fleischmann’s reputation was not going to make such a mistake, therefore Pons and Fleischmann must have had the evidence, but they had been rushed to publication by the

competition from BYU. Thus the paper was poor, but the work, should all the details be known, was not. To imagine that Pons and Fleischmann,

especially Fleischmann, had made these claims without doing the definitive experiments, as the paper seemed to indicate, was a stretch. The third possibility, which became accepted around the University of Utah as the gospel, was that Fleischmann and Pons must be hiding something. As Albagli said, ““When you get this incomplete information and you give them the benefit of the doubt, you realize they might be trying to buy themselves some time.” The history of science has a surprising number of precedents for this kind of covert business. When Galileo discovered the phases of Venus in December 1610, he claimed his discovery by anagram: ‘‘Haec immatura a me iam frustra leguntur o.y.’’ When unscrambled this becomes “‘Cynthiae figuras aemulatur Mater Amorum,’’ which means, with the poetic metaphors translated as well, that Venus emulates the phases of the moon. Galileo, as the sociologist of science Robert Merton put it, was a “seasoned campaigner’ when it came to defending his discoveries from claim jumpers. The precedent on everyone’s mind during cold fusion was Paul Chu’s 1987 discovery of a compound that remained superconducting at 90 degrees kelvin, which by superconducting standards was hot enough to be remarkable. It was the discovery of high-temperature superconductors that had momentarily made the scientific community willing to believe in miracles. Chu had submitted a paper in the winter of 1987 to Physical Review Letters claiming that the magical ingredient in his superconductor was ytterbium when in fact it was yttrium. Later Chu blamed the error on his secretary, who

had also, so Chu

said,

incorrectly transcribed the proportion of ytterbium, or yttrium, or whatever. (As one writer put it, “She must have been having a very bad day.”)?* Thus, when Stan Pons spoke at an American Chemical Society meeting in Dallas two weeks later, one suspicious individual

166 @ GARY

TAUBES

asked if there were any “typographical errors in the paper worth commenting on.” Hugo Rossi, who was dean of the College of Science at Utah, said he

began to have the opinion, or had been given the opinion, that Pons and Fleischmann had written their paper so that it would satisfy the requirements for establishing priority but give no more information. “Martin and Stan,”’ Rossi said, ‘‘did not want to get into a position of having to compete in research with millions of laboratories.” John Flynn, who was an expert on patent law at the University of Utah, remarked that he was called in by Chase Peterson after the announcement to help shut the barn door, so to speak, on the patent issue. He remembered advising Peterson and his colleagues that the less said about the technical sides of the discovery the better. “If this [had been] in a commercial lab,” Flynn said, ‘‘the place would have been locked and sealed and those people wouldn’t be allowed to talk to anybody.”’ Flynn observed that Pons was a canny businessman, familiar with inventions and patents, and he knew the risks of a public announcement. So it seemed eminently possible at the time that the paper had been purposely written to be obscure. There was, of course, an older historical precedent for this as well. Alchemists, traditionally, refused to disclose the secret of the philoso-

phers’ stone, which would transmute base metal into gold. As Artephius, an alchemist of moderate repute, put it, “Poor idiot! Could you be so simple-minded as to believe that we would teach you clearly and openly the greatest and most important secrets?’ Or as the good Reverend Charles Mackay put it in Extraordinary Popular Delusions and the Madness of Crowds: Unluckily for their own credit, all these gold-makers are in the same predicament; their great secret loses its worth most wonderfully in the telling, and therefore they keep it snugly to themselves. Perhaps they thought that, if every body could transmute metals, gold would be so plentiful that it would be no longer valuable, and that some new art would be requisite to transmute it back again into steel and iron. If so, society is much indebted to them for their forbearance.

In the end, the quality of the paper didn’t seem to matter. By the time it appeared, anyone who liked the sound of cold fusion was hooked already. It had become aself-sustaining phenomenon. Who, whether chemist or physicist, having tried the experiment and become caught up

BAD

SCIENCE

# 167

in the excitement, could read the paper, suspect it was scientifically bankrupt, and walk away?

24

CAMBRIDGE.

GAMMA

RAYS

Amid all the talk of heat and neutrons, it was the gamma rays that should have been the definitive proof of cold fusion. Pons and Fleischmann had placed their gamma ray spectrum in figure 1A of the paper, the position traditionally reserved for the strongest evidence inascientific case. Here, the author is saying, look, this is my proof. The reaction in question is known as neutron capture on proton. The

fusion event emits a neutron. The neutron is captured by the hydrogen in the water around the cell. This creates a deuterium nucleus, which

then releases a gamma ray with exactly 2.22 million electron volts (MeV) of energy. It always releases a gamma ray of exactly 2.22 MeV. The universe is that consistent. “It’s the first thing you get in first-year quantum mechanics,” explained Richard Petrasso, an MIT physicist who should know. ““They’ll talk about the deuteron binding energy, and it’s 2.22 MeV. It’s famous and been around for ages, and if you see a gamma ray coming out at 2.22, you know what’s taking place.” Counting

neutrons,

as Petrasso

put it, “can mean

diddly,” but a

gamma ray spectrum, that is an art form with a long and established history. Put a prism in front of a light source, as Isaac Newton first did, and the light fans out into its component colors. Put the prism in front of sunlight, and this spectrum is banded by hundreds of dark lines, each

of which represents the absorption of the light at exactly one wavelength by a single chemical element or isotope. If the line is there, it means the element is there. All spectra have valleys, which are absorption lines, and peaks, which are the signatures of distinct atomic processes that emit light, which is composed of photons or gamma rays, which are nothing more than high-energy photons. The same lines appear from the flame of a candle or the light of a galaxy across the universe. This is true in Utah, in China, on the moon, or anywhere else.

Petrasso explained this unequivocally: “Each atom has its particularly unique song; each gamma ray is a particular identification. It’s like your name, my name, anything. You showa spectrum, and that conveys a

168 8 GARY

TAUBES

tremendous amount of information and it’s very unambiguous. These are powerful signatures. . .. And when you show that line at 2.22, it means one thing, it means neutron capture on proton and it means that you’re generating neutrons.” And there it was in figure 1A of the paper. Gamma rays at 2.22 MeV (or 2220 KeV in the units preferred by Pons and Fleischmann), a peak that stood up like the Statue of Liberty or the Eiffel Tower. Unmistakable. Unambiguous. 25,000

20,000

|—

15,000

|—

COUNTS

10,000

|-

5000: j—

0

| 2,000

| 2,200 ENERGY / keV

| 2,400

Petrasso looked at that spectrum and found it “‘pretty startling.’ And then, as he continued to look at it, he said he grew very disturbed. That

a gamma ray spectrum can be disturbing may be the kind of thing that only a serious student of science can appreciate. Nonetheless, these gamma rays became one of the more unfortunate stories in the cold fusion episode. Petrasso was certainly not the only physicist who noticed that the gamma ray spectrum had its problems, but he had followed up to find out why. “If you look at my career,” he said, “you might summarize it by the following: if something makes X rays or gamma rays, I want to know how many, what kind of X rays or gamma rays, and what’s behind it.’’2” Talk to enough scientists, and the impression grows that they can be classified into two types: those who went into science because they

BAD

SCIENCE

§ 169

were good at it, and those who were obsessed with science and had no choice in the matter. Petrasso, at age forty-four, was of the latter variety. He was the son of two musicians, but he said he knew he would

be a scientist when he was five years old. Petrasso earned his Ph.D. at Brandeis University and studied, coincidentally, under Ryszard Gajewski. Over the years he had worked on the Uhum satellite, an orbiting X-ray observatory, and then on the X-ray telescope on the Skylab mission. He studied X-ray flares from the sun and for the past decade had been working on X-ray diagnostics for the MIT fusion program, developing techniques for analyzing the plasma in a tokamak by studying the X rays that it emits. Petrasso considered cold fusion immediately implausible and was unwilling to abandon his faith in nuclear physics on the strength of a press conference, which, he said, “‘is just glib bullshit, as far as we know.”’ He

did not join the Parker-Wrighton cold fusion collaboration because he did not believe cold fusion warranted that kind of effort. Even a year later Petrasso would grow exasperated when discussing cold fusion and especially gamma rays. “For me,” he said, ‘“‘the compelling piece of information is the gamma line. If they've got something, they’ve got to show me this line. And if they’ve got it, then, fine, I’ll go out and try and reproduce the experiment, but not beforehand.”’ Once Petrasso glanced at the Pons-Fleischmann gamma ray spectrum, he knew it had serious problems. For starters, the spectrum was clipped, which is to say it was not a spectrum but a single peak, identified at 2.2 MeV, with nothing on either side. It was like showing asingle line of color from the spectrum of visible light, except that gamma rays, which are beyond the visible spectrum, are therefore identified not by color but by energy. If Pons and Fleischmann had displayed their entire spectrum, then researchers could have judged for themselves that the gamma ray peak was at 2.22 MeV. Instead, they had to take it as a matter of faith, which was, of course, exasperating.

Why didn’t they show the entire spectrum? Petrasso didn’t know, but, he said, “‘in order to get 2.22 without showing the background lines just waves a flag in my face.” Even as an isolated gamma ray peak, the peak Pons and Fleischmann displayed was disturbing. It’s a scientific fact that all gamma ray peaks have certain identifying features, the same way faces or anything else have. In particular, they have what are called single and double escape peaks and Compton edges, all of which are properties of the way gamma rays interact in the detectors rather than of the gamma rays themselves. Suffice it to say that when the incoming gamma ray smacks the crystal

170 8 GARY

TAUBES

9

>

@

=

K~°

6

(1.460)



@

Dominantly

eo

gaBir'

=

(1.764)

38

aaBi2!4

3

salle

(2.615)

0 4

9g

1.4

leg

2.4

2a

Energy (MeV) in the detector, it will occasionally skid off to one side, or bounce backward out of the detector, or create new particles, even new gamma

rays that might escape from the detector unnoticed. Any of these would result in the detector failing to register some portion of the incoming energy of the gamma ray. Thus, gamma rays of a particular energy show up smeared across a range of lower energies as well. In this smear would be the single and double escape peaks, and the smear would end at the Compton edge. If these peaks and edges “‘ain’t there,” as Petrasso said, “it’s not a gamma ray.” These characteristics seemed to be lacking in Pons and Fleischmann’s single peak, although it was difficult to tell because they hadn’t displayed even enough of the background on the low-energy side of the peak to be sure. In the errata to the paper, which appeared shortly after, Pons and Fleischmann still neglected to show the entire spectrum, but they did replot the single peak, revealing slightly more of it on each side. Now it was an inescapable fact that the escape peaks and the Compton edge were not there. Thus, as Petrasso noted, it wasn’t a gamma ray peak. Once the errata appeared, it was possible to measure the width of the new peak. Petrasso did this, as did quite a few other physicists, and realized that the width of the peak was incompatible with the sodium iodide detector Pons and Fleischmann had used. Detectors are like cam-

BAD

SCIENCE

8 171

eras; they can achieve a certain resolution of detail, depending on their quality. The better the detector, the more precisely it can identify the energy of the gamma rays, and the narrower the peak. Pons and Fleischmann’s peak was impossibly narrow considering the instrument they had used. So it wasn’t a gamma ray peak. All these defects strongly suggested that the peak was what scientists call an artifact, a bogus signal produced by the equipment in question, which in this case would be the detector. Petrasso considered this an “ronclad”’ fact. This then raised the pivotal question: If the peak was an artifact, how did it come to have exactly 2.22 MeV of energy, precisely what would be expected for a sophisticated test tube generating neutrons? If the peak was an artifact, it could have manifested itself at any energy at all. Petrasso

found this coincidence very suspicious. ‘“What’s the chance of an artifact coming up at precisely the energy 2.22 MeV?” he asked. “It’s just about zero.” Lewis Thomas has written that, when scientists are presented with an

intellectual challenge, they respond with something very much like aggression: “While it is going on, it looks and feels like aggression: get at it, uncover it, bring it out, grab it it’s mine! It is like a primitive running hunt.” In this case, it didn’t matter that the challenge had nothing to do with adding another brick to the growing edifice of science. Petrasso was affected by the same powerfully hypnotic stimulant: get at it, uncover it. Find the answer. This was further stimulated, one

can assume, by his distaste for glib bullshit. ““We already knew that the data we were looking at was bogus and was an artifact,” he said, “‘and it’s

very unnerving because you have to ask yourself, What in the hell is going on here?”’

25

DEBRECEN, AND

TOKYO,

HUNGARY, JAPAN

Observations are useless until they have been interpreted. The analysis of experimental data therefore forms acritical stage in every scientific inquiry—a stage which has been responsible for most of the failures and fallacies of the past. E. BRIGHT WILSON, An Introduction to Scientific Research

172 © GARY

TAUBES

April Fools’ Day. Two Hungarian physicists, Gyula Csikai and Tibor Sztaricskai of Lajos Kossuth University at Debrecen, announced that they had duplicated the electrolysis experiment and detected the emission of neutrons. This was actually a confirmation, albeit premature, of

the low-level neutrons reported by Steve Jones and company. Jim Brophy called it the second confirmation of Pons and Fleischmann, intimat-

ing, of course, that the first was Jones. A Wall Street Journal correspondent interviewed the triumphant Hungarians and reported that “Mr. Sztaricskai was ambiguous about his results, at times saying he had confirmed the Pons-Fleischmann results, and at other times saying he had replicated the lesser, Brigham Young findings.” Paul Palmer of BYU was less ambiguous. The Hungarians, he said, “‘undoubtedly” confirmed the BYU room-temperature fusion discovery. In Tisbury a British reporter caught up with Martin Fleischmann returning home from his local pub and told him that their experiment had been successfully reproduced. Fleischmann responded, “‘Tf it’s true, that’s fine.’ It was not true, however. Mr. Sztaricskai and his confirma-

tion soon vanished into the background, never to be heard from again. Equally premature was the announcement by Professor Noboru Oyama of the Tokyo University of Agriculture and Chemistry. Nature reported that Oyama had detected “large amounts of heat and gamma rays’ generated by a cold fusion experiment. Apparently Oyama had performed a quick and dirty experiment so he could report results at the April 1 meeting of the Chemical Society of Japan. However, under such time constraints, if less dramatic explanations existed for the experimental results, Oyama had little time to root them out. Other Japanese researchers who had seen the data were not wildly impressed. The Nature report continued, ‘“Oyama says he will try to duplicate the experiment with the Japan Atomic Energy Research Institute to determine neutron yield.” This sounded like the kind of experimental procedure Oyama might have wished to pursue before reporting his results, but, in the brave new world of cold fusion research, the priorities had changed and therefore so had the rules.

BAD 26

SALT

LAKE

GROUND

SCIENCE

# 173

CITY.

ZERO

On April 4, Marvin Hawkins was fired. He found aletter in his school

mailbox from Stan Pons, who was lecturing on cold fusion at the University of Indiana. The letter was marked “confidential.” “So I rip it open,” said Hawkins, “and to my utter dismay—you could have picked me up with a putty knife off the floor, after having read this—it said he was surprised but relieved that I had returned the remainder of the stolen documents and materials to him. He was glad there didn’t have to be any other involvement by the law enforcement agencies or the universities or the lawyers, [but] he was disappointed in my behavior, and he thought it would be appropriate [if] I relinquish my ties with him and the university and return my key to the secretary. And I think, What am I reading here? [I] turn it over andI read it again, just to make sure it wasn’t a fantasy or some crazy dream. I’m going, ‘I can’t believe this.’ This is just unreal. This isn’t right. I had taken a lot, and I will not be scraped out of this place without a degree.”’ Hawkins took the letter to public relations and showed it to Barbara Shelley, who had been his contact in the cold fusion work. She seemed sympathetic, and he told her that if the situation was not rectified he’d go to the press with it. He made the same argument to Cheves Walling, the most distinguished chemist at the U and a member of the National Academy of Sciences. Walling said he’d talk to Pons; they’d work it out. Late that night, when it was already early morning in England, Hawkins called Fleischmann. He recounted the turn of events, and Fleischmann said, “Marvin, don’t worry about it. Don’t even talk to Stan. Let

me talk to Stan. .. . When Stan wants to talk to you, then that’s fine. Do not let us down. It’s going to make the situation worse.” On the morning of April 5, Pons tracked down Hawkins and they arranged to meet in his office, where they were joined by Walling and Shelley. Apparently, under some pressure, Pons chose to forget the whole thing and told Hawkins to get on with his thesis. Hawkins said one of the conditions he insisted upon at this meeting was that his name be added to the list of authors in the errata, which it was: ““M. Fleisch-

mann and S. Pons regret the inadvertent omission of the name of their co-author, Marvin Hawkins. . . .” Few scientists, however, believed that

M. Fleischmann and S. Pons had forgotten about Marvin Hawkins in their rush to get out the paper. As Hawkins later said with noticeable

174 B GARY

TAUBES

bitterness, “When somebody says, ‘Oops, we screwed up, we inadvertently omitted one of our coauthors’ names and we apologize,’ that’s absolute bull. Anybody would come across that would realize that’s the case.” In any case, later Cheves Walling told the story of Marvin Hawkins and the lab books to Chuck Martin of Texas A&M, pretty much as Hawkins told it. Martin, in fact, was visiting Pons on April 5, the day the lab book issue was finally resolved. Martin had hoped to get details of the experiment from Pons so he could successfully replicate it. “I was frustrated by the fact that I couldn’t get at him,” Martin said. ‘He looked like hell. Clearly he hadn’t slept in weeks. He was in some weird low gear that I’d never seen in him before, because he was so tired.” As Martin remembered it, Pons told him he couldn’t show him the

notebooks, because “‘this kid” had stolen them. This did not jibe with Hawkins’s version, in which the books were already returned by this time. Pons did give Martin a cursory tour of the lab and let him briefly glimpse the photocopies, then suggested they go to his house, where they sat and listened to messages on Pons’s answering machine. “He got phone calls from the Japanese consulate,” said Martin, “from Senator Al Gore; he got phone calls from congressmen, from quacks, from millionaires, from everyone.” Again, Martin tried to get Pons to talk about the data and failed. ‘Then the strangest thing happened,” Martin recalled. At Pons’s suggestion, they went off to a hotel, where Pons took to drinking at the bar with a television crew from the Canadian Broadcasting Corporation that was in town putting together a cold fusion documentary. Martin was dumbfounded:

“I’m sitting, thinking, Shit, we could be talking science

and instead we’re sitting at a bar drinking. I figured, Okay, he’s under a lot of pressure and maybe he needs this. But it was really strange. Here was Stan in the middle of the day, three or four o’clock, over here

drinking with these guys. He wasn’t talking science. He was talking to these guys from the CBC, and they were his chums.”’ Martin decided he was wasting his time and took an evening flight back to Texas.

BAD 27

THE

SCIENCE

§ 175

PRESS

On April 5, The Wall Street Journal reported that Brookhaven National Laboratory on Long Island had confirmed the BYU observation of neutrons from cold fusion cells. This was then considered additional confirmation of Pons and Fleischmann, as most of the press and public hadn’t differentiated between the two competing groups with their two competing claims. The Journal quoted Kelvin Lynn of Brookhaven saying, ‘“We’re not absolutely certain, but we have detected neutrons [produced by fusion reactions] that are consistent with the Brigham Young resules® This seemed to lift hopes for the folks back home in Utah. Chase Peterson told The Deseret News that Stan Pons had told him, “If I was so

sure two wecks ago, I am The News also reported ment with ordinary water nificant heat,” The News process is indeed nuclear

more sure now.”’ that Pons had finally done his control experireplacing heavy water. ‘‘It produced no sigsaid. ‘““This could be proof that the heating and not chemical as some physicists have

suggested.” Two days later the Associated Press reported that Brookhaven Laboratory had not, as previously reported, confirmed the production of neutrons in cold fusion cells. The Brookhaven scientists ran two experiments. One produced nothing, and the other turned up results so small as to be statistically meaningless. Then AP quoted Mark Wrighton of MIT saying that he and his colleagues remained very skeptical. “If nuclear fusion occurs,’’ Wrighton said, “‘it is at a very low level and our

detectors aren’t sensitive enough, or it takes longer than ten days, or it doesn’t work.” The story then turned to the University of Utah for comment and quoted Pam Fogle saying ‘“‘that the team believes the reaction can take ten days to three weeks to occur.” It had been only two weeks since March 23, so perhaps Fogle was suggesting that Wrighton be patient. On the other hand, nine days before Pons had told Al Bard that the reaction took 7.7 hours to occur.

176 BP GARY

TAUBES 28

SALT

LAKE

CITY

‘He that doeth nothing is damned, and I don’t want to be damned,”’ said

Governor Norman Bangerter, shepherding his flock into the promised land. With such apocalyptic oratory, the House and Senate of the state of Utah voted nearly as one—96 yea, 3 nay—to appropriate $5 million to the cause of cold fusion. ‘Waiting and seeing,” proclaimed W. Eugene Hansen, chairman of the state board of regents, “‘could mean that the discovery of the century will be developed by Mitsubishi. If there is to be a Fusion Valley, we feel that the Fusion Valley should be here, in the state of Utah.” The legislature proposed that the $5 million be split over two fiscal years, with $2 million for the first, and $3 million for the second. As The

Salt Lake Tribune reported, this cash would be spent for “equipment, space, retention of consultants, conferences and other ‘appropriate’ applied science techniques that might lead to market development.” The money would only be disbursed after the scientific community provided independent confirmation of the cold fusion findings. What form this independent confirmation would have to take was not specified. One can assume that on April 7 a convocation of lay politicians believed that such confirmation would be indisputable when it came. Yea or nay. A sign from the heavens. The governor himself would appoint a nine-member panel to decide whether confirmation had been achieved. The panel would consist of “‘a nuclear physicist, a chemist, an at-large ‘member of the scientific community,’ the state’s science adviser, a certified public accountant, two business-sector members with ‘tech-

nological research and development’ backgrounds and two at-large members of the general public.”’ And, finally, the legislature proclaimed that this fusion council would be a secret organization; should the people of the great state of Utah want to watch how their cold fusion money was being spent, they were out of luck. This was not, as one might think, a clever ruse by the Utah legislators to protect their reputations should they blunder mightily. Rather, as they later explained, they only wanted to assure that the valuable methods of cold fusion technology could be discussed without fearing that foreign companies would run off with them. Nonetheless, the local Society of Professional Journalists objected, pointing out that the state’s Open Meetings Law prohibited such secret business. The governor countered by signing an exemption to the law that would let

BAD

SCIENCE

#177

his fusion council close any meeting in which they expected to discuss intimate technological details. All budget sessions were open, all budget proposals would be public documents, but anyone caught leaking the technology would go to jail. The only cold fusion proponent in Utah who seemed to be against this kind of support was Stan Pons. Pons seemed to see all the cold fusion boosterism as a misguided effort to hang his reputation as far out on a limb as the citizens of his state would allow. In fact, he had begged Hugo Rossi to stop the $5 million from going through. “TI really didn’t understand Stan’s motivation,” Rossi recalled, “and I asked him about it, and

he said, “We don’t really know that this is going to work. This is completely premature to be doing this.’ ”’ Pons said that Fleischmann had mapped out a research program, complete with their estimated spending for the next eighteen months. He said the money promised from the Department of Energy and the Office of Naval Research would be enough to support them. Rossi responded that the rationale for the state funds was to keep their research ahead of the rest of the world, that ‘laboratories with millions

of dollars to spend were working on this, and in no time they would be way ahead.” But Pons disagreed. ““They can’t get way ahead of us,” he said. ‘““We have already put five years into it, we have learned lots of things, lot of subleties you have to work through. We'll be fine if we’re just left to do our program.” This argument to proceed conservatively had a great deal to recommend it. Pons, however, was no longer in charge. Once the legislature had allocated the $5 million, Pons was left to tell reporters that the state’s faith in cold fusion was a “‘great honor.” What else could he say? He added that now they would be able to proceed even faster with their research.

;

29

SANTA

BARBARA,

CALIFORNIA

Steve Koonin’s initial reaction to cold fusion was that it was crazy: ‘What are these guys talking about? This can’t be right.” But after he thought about it a moment, maybe it wasn’t so crazy: “‘I

178 B GARY

TAUBES

don’t know anything about palladium, but I know it absorbs a lot of hydrogen—I don’t know what ‘a lot’ means, either, [but] maybe in some funny way, some things can happen.” The question, of course, was What kinds of things? Koonin was on sabbatical at UC Santa Barbara when cold fusion broke. He was a professor oftheoretical physics at Caltech, where he also ran the Kellogg Laboratory, an experimental physics laboratory. He was a second-generation Lithuanian from Brooklyn and still had the pugnacious demeanor ofa New Yorker, looking not unlike an adult version of one of the Dead End Kids. He had done his undergraduate degree at Caltech in the usual four years, although he started when he was sixteen, and obtained his doctorate at MIT in three. He had been back at Caltech ever since. It was Koonin who had supervised the JASON review of Steve Jones’s muon-catalyzed fusion. Koonin initiated his cold fusion research from a neutral position. He perused the literature and simply collected facts. He learned, among other things, about the structure of palladium, the thermodynamics of the hydrogen-palladium system, the time it takes hydrogen to diffuse through a palladium lattice, and the energy required to absorb hydrogen (or deuterium) into the palladium. This last point was a crucial one. Pons and Fleischmann had claimed to get inexplicable amounts of energy out of their palladium rods. “So,” Koonin explained, ‘“‘you want to know how much energy do I really get out if I just dump all the hydrogen out again? Now that turns out to be a lot. You could explain an awful lot of heat on this deabsorption of hydrogen. There’s a lot of energy in there, equivalent to a reasonably high explosive.” Because Koonin had once written a textbook on computational physics, he then decided to use one of its programs to do “‘the most accurate calculations that one could ever get” on the possibility of two deuterium nuclei fusing. His partner in this endeavor into nuclear theory was Michael Nauenberg, a German-born physicist out of UC Santa Cruz. Nauenberg had grown up in Colombia after his family fled Germany in 1939. He had obtained his doctorate under Hans Bethe at Cormell and in 1966—after stints at the Institute for Advanced Study at Princeton, at Columbia, and at Stanford—had become one of the first faculty members to join the UC Santa Cruz physics department. Over the years Nauenberg had dabbled in condensed matter physics, statistical mechanics, astrophysics, nuclear physics, and dynamical systems, commonly known as chaos theory. Nauenberg was also recharging his batteries in Santa Barbara.

BAD

SCIENCE

8179

Step one in Koonin and Nauenberg’s theoretical approach was to establish a bench mark, what Nauenberg would call a fiducial point. This would be a simple situation that they could solve definitively, upon which everyone would agree, and from which they could proceed to more complicated and disputatious scenarios. Solving the fusion rate of deuterium nuclei in a palladium lattice would be very complicated, so they would begin with the spontaneous fusion rate of a deuterium molecule—simply two deuterium nuclei bound together by two electrons. This rate turned out, said Koonin, to be “ridiculously slow.”” Deute-

rium molecules would fuse spontaneously on the average of once every 10% seconds. Or, to put it a different way, given 10“ deuterium molecules, which is a lot of molecules, one of them might spontaneously fuse every second. “Or if you consider the sun,” said Nauenberg, “completely cold, but made up entirely of deuterium, there would be one fusion every year. That gives you a sense immediately of how crazy this really was.’’?? Each sophisticated test tube would contain approximately 10” pairs of deuterium nuclei. That was true for both the Utah and BYU variety test tubes. Thus, according to nuclear theory, both Jones and Pons and Fleischmann could expect one spontaneous fusion every 10%? seconds (10/107 = 10%) in their cold fusion cells. For Jones to observe two neutrons per hour, considering the inefficiency of his detector, his cells would have to be emitting roughly one per second. This was 10% times greater than the theory predicted. To produce the level of heat Pons and Fleischmann had observed—say, one watt of fusion power—the Utah cold fusion cells had to generate about a trillion, or 10’? fusion events per second, which was 10° times greater. So, as Koonin put it, “Jones had to buy forty orders of magnitude and Pons and Fleischmann had to buy more than fifty.” But then, he would add, “‘if you’re really an optimist, you say, ‘Look, if I could really get to forty, then what’s another ten.’ ’”° To say that buying forty to fifty orders of magnitude sounds like a lot is to make a Promethean

understatement.

(Although, for the sake of

pedagogical digression, these numbers are not so enormous that they don’t have names: forty orders of magnitude is ten duodecillion; fifty orders of magnitude is 100 quindecillion.)** Consider that the difference

between the diameter of a proton, which is the nucleus of a single hydrogen atom, and the diameter of the entire universe is only forty-one orders of magnitude. Or consider that the difference between the speed of light in a vacuum and the speed of a snail crawling across a country

180 #8 GARY

TAUBES

lane on a hot summer day is no more than twelve or maybe thirteen orders of magnitude, depending on the snail. Nonetheless, Koonin and Nauenberg went about calculating how those rates could be achieved. This, as it turned out, would require

compelling the deuterium nuclei within a palladium lattice to be separated by only one tenth their preferred distance when comfortably ensconced in a deuterium molecule. That means rather than sitting at 0.74 angstroms—an angstrom being 10 ~ '° meters—as they do in a molecule, they would have to be at 0.074 angstroms. It is one of those vagaries of quantum mechanics that one could get a 10 duodecillion increase in rate with only a factor of ten decrease in distance. This last reduction, however, was more

difficult than it sounds. A

factor of ten compression in one direction represents a factor of 1000 compression in volume, which is a proposition remarkably similar to squeezing blood from the proverbial turnip. Koonin and Nauenberg did consider one dim possibility: that this compression could be achieved if somehow the electrons in the neighborhood appeared to have ten times their conventional mass. This would have to happen by virtue of some mysterious property of the palladium lattice, heretofore undiscovered, as the aficionados of cold fusion liked to put it. This concept is known as the effective mass of the electron. And if these massive electrons existed, they would pull the deuterium nuclei to within the required distance, the same way muons do in muon-catalyzed fusion. Koonin and Nauenberg indeed checked their mathematics by plugging into their equations the mass of the muon—207 times that of the electron—and getting out the fusion rates that had previously been calculated and experimentally confirmed for muon-catalyzed fusion. Said Nauenberg, ‘““That was a nice check.” Physicists over the years had occasionally invoked the concept of effective mass to explain solid state phenomena, such as the electrical conductivity properties of semiconductors. But the concept is only meaningful as a “long-range effect.” In other words, the mass of the electron appears heavier because it seems to be distributed over many atoms inside the lattice. Cold fusion required that the electron swell into tenfold obesity within the orbit of a single molecule. This consideration launches the elephantine electron back into the realm of magic. Nauenberg pointed out that “short of saying all of quantum mechanics has to be thrown out the window,” it wouldn’t work. One relatively fruitful piece of physics did spring from the calculation, however. Koonin and Nauenberg discovered that fusion between deuterium and hydrogen would actually proceed more easily than that be-

BAD

SCIENCE

#181

tween deuterium and deuterium or even deuterium and tritium. This is because the rate of fusion is determined by an effect known as quantum tunneling, and the rate of quantum tunneling is determined by the mass of the nuclei involved. The lighter mass of the hydrogen atom increases the tunneling rate ‘““enormously,”’ which is to say from one every 10% seconds to one every 10°° seconds. ‘“‘So,’’ Koonin said, ““we made the

suggestion that [deuterium-hydrogen fusion] is going to go faster [than deuterium-deuterium fusion] and that people should try mixtures of protons and deuterons, mix light and heavy water.’ Koonin and Nauenberg wrote up their disheartening calculation and submitted it to Nature. They also typed it into the computer network, so that anyone with a terminal could see the evidence of the theoretical improbability of cold fusion. They mailed papers to all their friends. Pass it on, they said. But, once again, this was just theory, which is to say, it

didn’t make the newspapers. And it could certainly be ignored if it disagreed with experiments.

The second weekend of April was a glorious one in England, sunny, warm,

and green.

The Harwell

cold fusion team, however,

spent it

indoors looking for neutrons from a cold fusion formula of the sort advertised by Steve Jones. So they were sacrificing their weekend in the attempt to buy forty orders of magnitude rather than fifty. Dave Findlay, Martin

Sené, Derek

Craston, and Dave

Williams

took two identical

Jones-type cells, one powered and one unpowered, and shuttled them in and out of their neutron detector by hand every five minutes. This gave them the neutron background measurement from the unpowered cell at the nearly identical time and absolutely identical place as the test cell. It was another way of doing a control. In this case the idea was to rule out cosmic rays as a source of the neutrons Jones had reported. At first the Harwell scientists hadn’t bothered with the Jones cells, because they considered Jones’s electrolyte recipe, his Mother Earth soup, in Williams’s word, “‘ridiculous.” After observing no neutrons with a variety of more

electrochemically

sensible concoctions,

they

decided that they would have to duplicate Steve Jones’s recipe to the

182 #8 GARY

TAUBES

letter or no one would believe their results. “It was Chris Jones,’’ said Williams, “who made this electrolyte up. He came over quite late and said, ‘I am ashamed to show you this.’ This little vial of black muck. So

we did that, and there was still nothing.’’* That the Harwell group was now exceptionally motivated to find evidence of cold fusion cannot be doubted, although Fleischmann would later claim it could. A rumor had wafted through Harwell that physicists at the University of Birmingham had successfully replicated the Jones experiment. This had provoked panic among the Harwell management, who started clucking about why Birmingham could do it and Harwell couldn’t. If cold fusion was real, said Findlay, ‘‘and it is going to be such an important technological thing, then the UK Atomic Energy Authority is going to look a bit stupid if they can’t find anything.” From this, one could deduce two possible conclusions. Perhaps there would never be neutrons and, ipso facto, cold fusion, at least the BYU

variety, was a canard. Or, the United Kingdom’s Atomic Energy Authority, being composed of mortals, after all, was still doing something

wrong.

31

COLLEGE

STATION.

PASADENA

On Friday evening, April 7, Chuck Martin and his wife, Deborah Greco,

a veterinarian at Texas A&M, began the two-hour drive from College Station to Austin, where the pleasure of their company was expected at a birthday party. They had been driving for about twenty minutes when Martin began to repeat like a mantra, “‘I can’t leave; I can’t leave; I can’t

leave. I’ve got to go back.” So they turned around and Martin went back to his laboratory, where earlier that evening one of his cold fusion cells had apparently begun to generate a considerable amount of excess heat. This seemed to be exactly what Pons and Fleischmann had advertised. Both of them had said look for heat, not for radiation, and now Martin was getting heat.

Martin was one of the better young electrochemists around. He was thirty-four years old, aggressive, and looked like a young James Garner. He had gone to a small college in Kentucky, where he had planned to study English but became a chemist, he said, because he couldn’t get

BAD

SCIENCE

§ 183

anything above a C on his English compositions, and chemistry came

easily. “So,” Martin recalled, “‘I thought I’d show these bastards, I’m

going to switch majors.”’ Martin went to the University of Arizona for his doctorate because it had a very good program in analytical chemistry, and because he was ready to see something different and it was a long way from his midwestern roots. Martin did his postdoc with Al Bard before coming to A&M.

Bard,

who had instructed some seventy postdoctoral fellows over the years, later said that if he would have had to guess which two would get positive results in cold fusion, Martin

would

have been one

of them. Bard

described Martin as “‘terrifically aggressive, very smart, but driven.” He said it as though it could be dangerous in science to have too much ambition.

Nate Lewis said that when

cold fusion broke, he made

a

mental list of the chemists he would want to review his results before he published. This list included his former mentor Mark Wrighton, whom he called “definitely a pro,” Barry Miller of AT&T, Larry Faulkner, Al Bard, and Martin. “If Chuck had come up with something convincing,” Lewis said, “‘we would worry about it.”’ When cold fusion came along, Martin, like Bard, went from his desk,

where he had managed his research group, back into the lab. He hadn’t done any research with his own hands in five years. Why not? “Because I had graduate students to do that,”’ Martin explained. As for cold fusion: “IT was in the laboratory twenty hours a day watching the experiment, making the adjustments, washing glassware. It was crazy. Every cell that went into that calorimeter, I put together with my own hands. Every one.” Martin had two good reasons for believing that he could do a better job at replicating cold fusion than anyone else in the world. The foremost was that one of his closest friends was Stan Pons. He had met Pons through the Office of Naval Research, which supported them both. (After the March 23 announcement, it was Martin who would tell Newsweek that “‘the essence of genius is taking an idea which some people think is ludicrous and seeing the possibilities.” He believed that applied to both Pons and Fleischmann.) When Pons and Fleischmann went public, the funding agents at ONR told Martin to spend whatever he needed to replicate the experiment. They would take care of his expenses. Martin’s second reason was Ken Marsh and Bruce Gammon, whom

he had contacted at the A&M Thermodynamics Center on the day after the announcement. Marsh was one of a dozen people in the country who could be called an expert calorimetrist. He was from Australia and had

184 8 GARY

TAUBES

been doing calorimetry for twenty-eight years. His partner, Gammon, a chain-smoking cowboy from Oklahoma, had been doing calorimetry since the 1960s. Marsh and Gammon’s calorimetric expertise was well known in chemistry. Paul Palmer, for one, said that when he asked one

BYU chemist why he wouldn’t do cold fusion calorimetry, the chemist replied that if Marsh and Gammon were doing it, they would get the right answer, so why bother? Martin hoped that Marsh and Gammon would tell him how to build a calorimeter and then he would go off and build one, but they were equally aware of the opportunity at hand. The A&M Thermodynamics Center was part of the Texas Engineering Experiment Station, TEES, and Marsh and Gammon’s jobs depended on obtaining contract work from industry and government. Business had not been great lately; in fact, there was some talk around the university of closing TEES down. Here was a discovery that depended, so Pons and Fleischmann insisted,

directly on their kind of expertise. They told Martin that they wanted in and suggested that they do the experiment together at the Thermodynamics Center. Martin agreed. When Martin received a copy of the Pons-Fleischmann paper, he brought it to Gammon and Marsh to analyze. Gammon considered the paper “consistent with what you would see for an undergraduate physical chemistry experiment.’ He and Marsh speculated on a number of areas where small errors might have crept into Pons and Fleischmann’s calorimetry, but their results still seemed so large that the errors didn’t seem relevant. Martin, Marsh, and Gammon decided to build the best possible

calorimeter and take their chances. Just what made calorimetry so difficult was the nature of the cold fusion experiment. Chemists, if they do any calorimetry at all, usually measure the heat released in a chemical reaction. This is a variation on a high school experiment, which is what makes it look so easy. Mix some chemicals together in a tin can, dunk it in a bath, and use a thermometer

to measure how hot it gets. What the chemist is measuring, however, is the total change in temperature caused by an isolated chemical reaction in a closed system: say the temperature of the system when the reaction starts is twenty degrees and shoots up to eighty degrees by the time the reaction 1s complete. The difference in temperature indicates how much heat was generated, which can then be converted to how much energy or power was generated. But that procedure is for a reaction that generates a certain amount of heat and then stops. What about a reaction that continues to produce heat

BAD

SCIENCE

@ 185

indefinitely, which is what the sophisticated test tubes were advertised to do? Pons and Fleischmann’s electrolytic cells, whether they generated sustained nuclear fusion or not, worked

like electric heaters. Put an

electric heater in a cup of water: the water gets hotter and hotter, and eventually boils. The temperature of the water is no indication that the heater itself is generating more energy than it’s taking in. In cold fusion, the rate at which the temperature changes has to be measured, and this is exasperatingly difficult. What Pons and Fleischmann had done was take the conspicuous shortcut. Rather than measure the rate of change of the temperature in their cells, which involved entirely too many unknown variables, they ran what was in effect a steady-state experiment. They let their fusion cells radiate heat into a surrounding water bath, then assumed that the temperature difference between the cell and the bath would be proportional to the amount of power produced by the cell. This required measuring the difference in two temperatures and converting it to power, which was also not as easy as it seems. If any heat managed to escape from the cell without raising the temperature of the water bath, that would affect the validity of the initial assumption. The power produced by the cells would no longer be proportional to the temperature difference between the bath and the cell. The same would unfortunately be true if any heat entered the cell or the water bath from somewhere other than the electrodes in the cell. And if the rate at which the cell radiated heat happened to be changing with time, that would also affect the validity of the assumption. It was impossible to judge from their JEAC paper whether Pons and Fleischmann had taken any of these variables into account, but Marsh and Gammon had

too many years’ experience in calorimetry to ignore them. Gammon considered good calorimetry to be, more than anything, a question of accounting. ““You draw some arbitrary boundary around the cell,” he explained, ‘‘and then you have to make sure that you account for all the energy and materials that pass through that boundary. That’s really all there is to it. And one just has to be a good accountant and be alert looking for all these things. You have to know what might be moving across the boundary. Temperature gradients, for instance. If you have temperature gradients, do they cause anything to move across the boundary? If they do, you have to eliminate them. If you’ve got materials moving across the boundary, are they taking energy out or in? Or are you moving materials—a gas, for instance—out in one place and drawing

186 # GARY

TAUBES

cold gas back in another place? Then how much heat is transferred across the boundary? You have one temperature on one side of the boundary and another temperature on the other. But does the heat transfer remain constant? There are a number of things that can cause that not to be true.” Pons and Fleischmann had made this accounting business more difficult by using a cell that was open to the air, which meant there was at least one route through which anything could enter the cell or leave it undetected. When Martin visited Pons in Utah on April 5, he noticed that the heat transfer in Pons’s sophisticated test tubes seemed to go up and down with time. Martin called it saw-toothed: as the heavy water in the cells was converted to deuterium and oxygen and the level of heavy water dropped, the rate at which the heat escaped from the cell steadily increased. Then someone would fill the cell to the brim with heavy water, and the rate would go back to normal. Martin asked Pons about it, and Pons didn’t have an explanation. He said, ““We don’t worry about it.” But Martin later realized that Pons and Fleischmann should have worried about it because it turned out to be one of several artifacts that most likely accounted for their excess heat observations. The alternative to an open cell, of course, was closing it off to the

outside world. This would make for easier accounting. But the cell was releasing hydrogen and oxygen and generating heat, and there was more than a little danger involved. (And, after all, Pons and Fleischmann had

already reported one explosion, although they attributed it to a nuclear reaction.) Pons and Fleischmann also insisted that the geometry of their cells was important to producing fusion. Just why the shape of a test tube might help induce a sustained nuclear fusion reaction was not clear. But anyone who wanted to replicate the experiment had to try to replicate the geometry of the cells. So open cells it was. It doesn’t take a great deal of experience to put together a more than adequate calorimeter, at least according to Gammon. “You don’t have to control the temperature to a ten thousandth of a degree,’ he said. ‘You don’t have to have all that precision. The problem is, when you get sloppy, you overlook things you should know better than to overlook.”’ The calorimeter Martin, Marsh, and Gammon built seemed sim-

ple and foolproof, although one could argue that nothing is foolproof on the first go-around. They submerged a Pons-Fleischmann-style fusion cell in a tank of water. The cell had the usual two electrodes, as well as

a heater for calibration. They would initiate each experiment by running a given amount of

BAD

SCIENCE

8 187

electric power into the heater, then monitoring the temperature rise in the cell. For example, running two watts of power into the heater might eventually produce a five-degree temperature rise. If they left the heater

on, the cell temperature would increase five degrees, a low simmer, and

stay there. Then they would turn on the current to the electrodes, which would also dissipate power and heat the cell. As Martin told it, if they ran one watt of power through the electrodes, and the electrodes were not doing anything peculiar, such as generating sustained nuclear fusion, they would then have to turn down the heater by one watt to keep the temperature steady. That was how they calibrated the system. “The amount we have to back off the heater,”’ Martin said, ‘‘is by definition

exactly equivalent to the power provided by the electrochemistry.” Okay, so what would happen if they put one watt of power into the electrochemistry and had to turn down the heater by two watts to keep the temperature steady? That would be excess heat. Somewhere, somehow, one watt of power would be stealing into the cell. Sustained nuclear fusion? Maybe. It could also be the result of an instrumental artifact, which would be what they’d have to find out. So Martin, Marsh, and Gammon began running experiments in the thermodynamics lab at A&M. They took turns baby-sitting the calorimeter and its cell, twenty-four hours a day, watching a strip of chart paper which recorded the temperature. As the temperature started to rise, they’d have to back off the heater a little. The temperature would fall, and they’d crank the heater upalittle. “In a situation where you don’t have excess heat,” Martin said, ‘‘you’d have a wavy line that would go up and down as you kept adjusting this thing to keep it in thermal balance. And we literally did it around the clock.” In the first two weeks of cold fusion, Martin talked daily to both Pons and Nate Lewis. He and Lewis often spent as much as three hours a day on the phone; however, it doesn’t come as a surprise that once Martin began observing excess heat on the night of the seventh, he told Pons but didn’t bother to inform his close friend at Caltech. Lewis’s researchers had not seen even a glimmer of excess heat. And he learned while calling around the chemistry circuit that no one else had seen it or knew anyone who had. Martin knew this as well and could take it two ways. Either he had a mistake in the making, or he was going to be first. The American Chemical Society had scheduled a special cold fusion session for their annual conference on April 12 in Dallas. It was only a three-hour drive from the A&M campus, so Martin wanted to get

a definitive answer by then if at all possible. This special session was being

188 @ GARY

TAUBES

who was head advertised as the Woodstock of cold fusion. Robert Park,

hed a of the American Physical Society’s Washington office and publis cold weekly newsletter on the computer network, called it “‘the first fusion shoot-out.” something Lewis had no plans to go to Dallas, unless, of course, he saw

definitive before then. On Saturday, April 8, John Gladysz, the Utah

on in chemist who was on sabbatical at Caltech, described Lewis’s situati

an e-mail message back to Pons: Stan, I've had a real ringside seat on the fusion efforts here at Caltech, and have filed several reports with my colleagues on the local drama. | haven't wanted to bother you until | had something substantial, because I’m sure you have had a zillion things to try fo do. Anyway, | just got back from 2 days at UCSD and Nate cornered me real quick because the scenario that he believes will take place at Dallas is that about 12 research groups will report that [they] haven't been able to match your heat or neutron yields. Nate is in the same boat at the moment, but knows his efforts were far more careful than some who claim no heat/no neutrons. Anyway the word he wanted to pass on, and| fully [corroborate], is that he’s still willing (and his students are still enthusiastic) about putting more work in to duplicate the experiment if you want to give more details. They are not publicity hounds (they have avoided all press contact) and are not seeking to compete in this area head-to-head or to in anyway violate a trust. So perhaps you have someone to speak up for you at the Dallas meeting. The thing | would vouch for is that if you think you need an ally, etc., the situation here is good for that. My understanding is that there is a special Weds session at Dallas on this fusion stuff. If these guys could get either heat or neutrons by late Tues night, they would fly out fo report positive results at the meeting (and would be happy to do so). Anyway, | trust your stewardship of this situation, and I’m not trying to give ‘‘advice’’; just the facts of

what is available on this end. I'm really excited by your discovery, keep pressing on. Bye, John. The same day Lewis sent Pons a lengthy account of what they’d done and their failure to observe cold fusion. It began:

| am holding to my promise not to talk to press, etc. about our work and to give it our best shot before saying anything to anyone else. Here is an update of our progress (or lack thereof) to date. We haven't been very successful to date, so | have included an extremely detailed description of our methods and procedures so that you can help us fix what we are doing wrong. He went on to describe their experiment in excruciating detail. Lewis said later he did it because he had no idea what was important: “We told

BAD

SCIENCE

# 189

him everything we knew. We still believed there was something missing.’”** And Lewis closed with aplea for help:

Obviously, we are missing a key step somewhere. . . . We just haven't yet seen any of the effects that you have reported. We will work like crazy trying to improve our mistakes if you can offer us any advice for changing the experimental procedure. We will keep our current cells running and check them periodically until we find something or hear from you regarding possible changes in the method. If they turn on, | guarantee that | will let you know immediately. Thanks for your help, and | hope that we can be of help to answer the doubters regarding the validity of these observations. Pons never replied to either letter. Perhaps he was so inundated with requests, calls, demands, and anxieties that he simply glanced at the letters, and, as Martin later speculated, “probably thought, Goddamn it

why can’t they do this experiment right?” Saturday night, April 8, Martin, Gammon, and Marsh wrote upa short paper to send to the Journal of Electroanalytical Chemistry. Martin had been in the lab since he had returned from the aborted trip to Austin. The cold fusion cell was still generating excess heat and had shown no signs of petering out. “That was the surprising thing,” Martin recalled, “because we thought initially that there must be something wrong, and we thought that it would stop. We ran it continuously before we finally thought, Well, it’s not changing. We recalibrated the system. We turned it off, and turned the heater back on to see if it took the same power to maintain that temperature that it did before.” When Martin told Pons what they had, Pons was elated. He began encouraging Martin to go public. Several times over the weekend, Pons assured Martin that he knew of research in the government labs that confirmed the excess heat as well, but that they were prohibited from going public. In the early hours of Sunday morning, Martin called Ron Fawcett, the

American editor of JEAC, at home in Davis, California, told him what was happening, then faxed him a copy of the paper. As Martin recalled it, “We said all right, we have the paper submitted. So what do we do? We knew that everybody else was doing this work. We heard dozens of rumors that this guy had this, this guy had that, somebody had heat, somebody had neutrons. We thought, We have this result; it’s an important result; we’re going to tell the world about it. Pons and Fleischmann had suspended all the rules. Everyone else is going to do it anyway, so let’s do it. We were convinced that there were twenty labs that could do

190 8 GARY

TAUBES

, the same. There wasn’t a single person who, if they had positive results wouldn’t have done the same thing that we did.” Martin called the A&M public relations people Sunday morning around nine and found the office open and ready to do business. He explained that they had a confirmation of cold fusion and they wanted to have a press conference the following morning. The Amierican Chemical Society meeting would be starting in Dallas, and, if they got the word out quickly, all the reporters covering the meeting could easily catch a plane down to A&M for the announcement. (‘And they did,” said Martin. “‘All of them.’’) Martin then called Kenneth Hall, who was deputy director of TEES and oversaw the Thermodynamic Center, in which they were running the experiments. Hall’s primary concern was that they contact the nuclear safety office and make sure they were not doing anything dangerous. Sunday afternoon the nuclear safety personnel checked the laboratory for radiation. They found none, but Martin did not consider that a bad sign. Pons and Fleischmann had no radiation either. By Sunday evening, Martin, Gammon, and Marsh began running light water control experiments. Martin also called Nate Lewis for the first time since they had started seeing the excess heat. “Well, what have you got?’ Lewis asked. And they discussed the details of Martin’s experiment, which seemed okay. The conversation, as Lewis reconstructed it, went like this:

“Chuck,” Lewis said, ‘‘are you sure this is right?” “I couldn’t have done it without Marsh and Gammon,” Martin said.

“Those guys are pros. We have the best calorimetry setup. We know this is right.” “Well, Chuck, we don’t see excess heat,” said Lewis. “Well,” Martin said, “I’ve got a press conference already scheduled.”

“Chuck, I don’t care about a press conference. You’re not first anyway. What does it matter?” “No, we got it,” insisted Martin. ““We’ve got to go out and do this.” “You don’t have to do anything,” said Lewis. “Look, let’s find out if this is right. Remember, we agreed on day one.” “No, I’ve got to go through with it,”’ said Martin. “The press conference is nine o’clock tomorrow morning.”’ “Chuck, how many controls have you done?” “I’m doing one now.” “Chuck,” Lewis said, ‘‘don’t have this press conference until you have

controls done. You’ve got to promise me that you’re gonna stay up all

BAD

SCIENCE

#191

night and that we’re gonna talk again, and before this press conference, youre going to tell me the results of these controls.” “I promise you.” “I’m going to call you if you don’t call me,”’ Lewis said, ‘‘and we’re going to find out. What time is the press conference?” “Nine A.M.” “Texas time?” hexas times “Okay,”’ Lewis said, “‘that’s seven A.M. my time. I’m going to call you at three A.M. my time if you don’t call me before then and tell me that these controls have worked.” Martin didn’t listen to Lewis. ‘‘At that point,” Martin later explained, “I believed Stan was right, and I believed no one else had enough details to do the experiment. I believed there was a recipe that if you followed it, it worked. And in my mind Nate Lewis is a great scientist, and if he doesn’t know the recipe he can’t make chicken cordon bleu either. I believed that I had a closer link with Stan, and I had information not

available [to anyone else]. And I knew I had good calorimetrists working with me. AndI knew no one else did. I knew that Nate didn’t know how to do calorimetry. He learned that from me. And, at the end, the effect

was just mind-boggling. It was a big effect. I didn’t listen to Nate.” Martin, Marsh, and Gammon

did the controls. It took several hours

to change the palladium rods, and after several hours the light water controls began to generate excess heat as well. Said Martin, “I said, ‘Hoo,

no, wait a minute. Light water response. This is starting to get screwy.’ ” Martin did not call Lewis. Instead, he called Pons, which constituted

a serious error in judgment, under the circumstances: ‘‘So I call up Stan. And I say, ‘Stan, we have a positive result in [light] water. Have you ever gotten that?’ He said to me, “Oh, Chuck, you are into the most exciting

thing now. This cold fusion works in light water, too. The natural [fusion rate] is low, but it’s high enough that you can get this effect to go in light water and we have seen it too. But we aren’t allowed to make this result public.’ I said, “Why not?’ He said, ‘I’ve got some secret agreements with the Department of Defense. I can’t report that result, but if you want to, you can, and it’s your baby.’ I go back to Marsh and Gammon, andIsay, ‘Jesus, he says he’s got a light water response also.’ And so we’re all excited. We’re going, ‘Jeez Louise, man, maybe we are on to something here.’ Because he says that he sees it, you know, and I’m believing this guy because he’s my friend, and because he’s an established scientist.’

192 © GARY

TAUBES

When Martin did not call Lewis back, Lewis called him at three in

the morning as promised. Martin told him that the controls weren’t working. Lewis said, “You mean they’re not giving heat?” Martin said, “No, we just can’t get them to work. We just can’t get the controls to work.” Lewis: ‘Call off the press conference. No one is going to blink if you wait a day later. It’s not going to be a big deal. Chuck, you promised me.” And, because this was cold fusion, Martin ignored him again. In fact,

the conversation may have been irrelevant bécause news of the A&M confirmation had already been given, by Pons apparently, to The Deseret News, which ran the story in the Monday morning edition.

Considering the portentous circumstances, the press conference went off without a hitch. Chuck Martin, dressed smartly in a suit and tie, showed

the reporters the experiment and read a prepared statement. He announced that Texas A&M had confirmed the observation that “when D,O is electrolyzed at a palladium electrode, more energy comes out of

the electrochemical cell than is put into the cell . . . in agreement with the findings of Pons and Fleischmann.” Martin was careful not to say he had confirmed cold fusion, only that their results “indicate it is worth-

while pursuing the possibility of electrolysis of heavy water as an energy source.” He explained that what they had seen could still be the result of some unusual chemical process, which they had yet to determine. When a reporter asked whether or not they’d done the light water controls, he said only that they had, which was true, however misleading. Martin found the press conference surprisingly easy to handle. It was just another talk, and he was accustomed to giving talks. “The enormity of it didn’t really hit me,” he said, “until people started calling me from London, from all over the world, saying they saw me in the paper.’’*°

32

ATLANTA,

GEORGIA

Hours after Chuck Martin threw his hat into the ring, or hurled himself into the abyss, depending on one’s level of optimism, James Mahaftey

BAD

SCIENCE

#193

and his colleagues at the Georgia Tech Research Institute did the same. Martin had been right in believing that other impetuous cold fusion researchers were ready to make the big leap. He had been wrong, however, in that most of them, like Mahaffey’s researchers, would still

be reporting the detection of neutrons. Mahaffey and Martin together would almost convince the world that cold fusion was a reality. When cold fusion broke, Mahaffey was thirty-eight years old, had a doctorate in nuclear engineering, and held the august title of senior research scientist at the Georgia Tech Research Institute. The research institute, however, was not the Georgia Institute of Technology, as one

would have assumed from the news reports of Mahaffey’s announce-

ment, but a research arm of Georgia Tech, which, like TEES at Texas

A&M, did contract research and development for the government and industry. Mahaffey himself had been working on a sophisticated computer programming project at a local air force base. This had nothing to do with nuclear engineering, but GIRI worked from contract to contract. ““We are doing computer analysis,’’ Mahaffey said, “because computer analysis sells. There’s no money in nuclear work. We’d like to do more, but we'll do anything anybody will pay for.” On the Monday following the Utah announcement, Mahaffey and his colleagues put together a quick and dirty cold fusion experiment, using a standard electrolyte, a bottle of heavy water, which they found in the lab, and a palladium ingot, which they bought from a rare-metals broker. “Tt was funny,’ Mahaffey said. ““We called up this rare-metals broker. She said, “You know, for years and years palladium has remained pretty much static, but just today it’s gone up 20 percent.’ That was our first indication that we weren’t the only curious people in world.” They borrowed a neutron detector and a Geiger counter, put the juice to the cell, and generated absolutely nothing. “‘Stone-cold dead,” said Mahaffey. This negative result, however, only made cold fusion all the more seductive. If cold fusion had been as simple as it appeared, they figured, it would have been discovered years ago. Perversely, this was a positive sign. Shortly after the first negative go-around, the administration at GTRI agreed to spend $25,000 on cold fusion research, which allowed Mahaffey and friends to do the experiment right. “We were very skeptical,” Mahaffey observed, “but Lord, we had to know, and we were getting zero out of Pons and Fleischmann.” Like Martin’s group, the GTRI researchers also believed they had two advantages over the competing multitudes of the curious. The first was Billy Livesay, a local materials scientist whose expertise was hydrogen in

194 B GARY

TAUBES

a metals and who said that it wasn’t off the wall to expect fusion in hydride. This may have been naive from the point of view of a nuclear physicist, but this expert was not an expert in nuclear physics. The second advantage, trivial as it may sound, was their choice of glassware. They speculated that Pons and Fleischmann had not observed many neutrons because they were using Pyrex glass, which contains boron, an element that is exceptionally good at soaking up neutrons. “We thought,”’ said Mahaffey, ‘‘that’s why they’re having such trouble with neutrons, and that’s why they’re so cagey about it.” So the GIRI crew used glassware of pure quartz, through which neutrons would pass readily. That was their ace in the hole. By Friday night, April 7, they had the experiment duplicated to the best of their knowledge. Like everyone else, they had received the Utah paper and sifted through it for pertinent information. They had four ancient but working neutron detectors, borrowed from the School of Nuclear Engineering, and blocks of paraffin to shield them should the cell work. One hour after midnight, Rick Steenblik, who had the second cold

fusion shift, called Mahaffey and said they had a neutron signal, which sounded suspicious already because Pons and Fleischmann had said it would take weeks. Mahaffey suggested that Steenblik double-check the neutron counters and apparently went back to sleep. Saturday morning, Mahaffey saw the data. “By God,” he said, “it was way above background. It looked like a neutron-generating incidence.” The nuclear reactor at Georgia Tech wasn’t running at three in the morning, so it could not have generated the neutrons. Mahaffey still refused to believe it. He suggested they run a test. He removed one of the detectors from around the cell and put it behind the boron blocks. If anything in the cell was generating neutrons, then they would not get through that boron to the detector. So they put the detector behind the boron blocks and, as Mahaffey put it, ““Boom!”’ the signal vanished. He removed the detector from behind the boron and placed it near the cell; the signal reappeared. ‘We did that all day,” he said, ‘‘one hour in, one hour out. And all

day long, when it was behind the boron blocks, nothing but background, and when it was near the cell, it was getting neutrons, hundreds per hour.

Not many—we’re used to millions per second—but a neutron is a neutron. If you’re making a neutron by putting a low voltage through a glass of water, something’s happening. We were very excited.”” They continued the same routine through Sunday, and continued getting the

BAD

SCIENCE

8 195

same results. They even called in a few experienced nuclear engineers, who seemed equally intoxicated by the data. Mahaffey on the next decision sounds eerily familiar: ‘“We started to think, What can we do now? We’d heard rumors that A&M was about to announce, and rumors that Brookhaven was about to announce, and

the week before the Hungarians had announced neutrons. We thought, Well, just sitting on this isn’t going to do anybody any good.” On Monday morning, April 10, Mahaffey and Livesay went to the administration at GTRI, in particular Jim Stetson, the vice president of research. They laid out the data for Stetson and didn’t say a word. Stetson took one look and called research communications, apparently the local euphemism for public relations. The GTRI research communications specialists then wrote up a twopage press release, showed it to Mahaffey and his colleagues, who thought it was fine, and sent it out on the wire. Mahaffey figured that was that, which, of course, it wasn’t. ““Well, this thing went over like the

nuclear weapon of press releases. Like nothing I’ve ever seen. All of a sudden, we were up to our necks in cameras. We weren’t prepared for that. Who would have thought they’d be so interested in what was really a very tiny experiment?” That was just the beginning. Immediately after the press conference, the GTRI neutrons joined Martin’s excess heat confirmation as the cold fusion news of the day, which is to say they were on the radio, CNN, the evening news, and by the next morning in every newspaper and, more often than not, on the front page. Mahaffey’s phone lines quickly jammed or just locked up. His secretary conceded defeat and went home early. Of the callers who did get through, Mahaffey talked only to researchers at the national laboratories—Brookhaven, Alamos, and so on.

Oak Ridge, Los

‘Immediately I had a real sick feeling,” Mahaffey recalled, ‘“because nobody else was finding neutrons, and they wanted to know what we had done to get our neutrons. I thought, We may be good, but we’re not that good. I thought everybody was on the verge of announcing, but nobody was getting anything. They were frantic for me to tell them what we’d done. And I didn’t leave anything out: I told them all about the quartz, what voltage we were using, everything. They were just frantic trying to copy us. ‘Damn, Georgia Tech has beaten us. Georgia Tech 1s doing something right.’ I immediately felt this real sick feeling. Something’s wrong.”

196

GARY

TAUBES 33

COLLEGE

STATION

Twelve hours after the A&M press conference, Chuck Martin and his colleagues also knew that something was very wrong with what they had done. Frank Cheng, a postdoc with Martin, described Monday evening as ‘“‘a disaster.” Martin didn’t like to talk about what had occurred between Monday night and Wednesday, which is understandable. He said later that he didn’t even like to think about it anymore. Cheng recalled the details, as did Del Lawson, a graduate student in

Martin’s lab. Cheng had been unnerved ever since they had run the light water controls Sunday night and seen excess heat. Cheng was twentynine and had come to study with Martin because Martin had a reputation as a chemist with imagination. Cheng expected to publish a lot of papers if he spent a year or two with Martin. Once cold fusion set in, that was not to happen. Monday morning, before the press conference, Cheng had mentioned to Ken Marsh that they had observed excess heat in the light water control. ‘“‘Doesn’t this cast some doubt on the results?’ Cheng asked. Marsh said he’d fill him in after the press conference and suggested that for the time being Cheng not mention it. Lawson was also anxious about the light water result and was told the same thing. Don’t mention the light water tests in public. He said, “We were just going to report the heavy water result because it replicated Pons and Fleischmann and then come out with our own more systematic study on the light water. We felt we had replicated Pons’s work, and that’s what Chuck [Martin] felt was important on that Monday morning.” These irrational reactions seem like temporary insanity, which may be as good a diagnosis as any. None of them had slept in days. Cheng called it “an interesting time.” Later that day Cheng heard about the phone conversation between Martin and Pons the night before. ‘‘[Pons] had said that this was an expected effect and therefore what we were seeing was some kind of a new energy source from light water, and this was supposed to be a secret, and there was going to be a Nobel Prize in all this. I thought I was standing in the lab that was going to give the energy source for thousands of years to come.” Monday evening Martin had them running controls, replacing the palladium electrode with electrodes of different metals. The idea was that

BAD

SCIENCE

8 197

if palladium was needed to induce fusion, jettison the palladium and see what the cell does without it. They put in successive electrodes of carbon, tungsten, then gold. All gave identical amounts of excess heat. “That’s when

it set in to all of us,” Cheng said, ‘“‘that this was

an

experiment that was in error. We did the control experiments after the press conference. This was the kind of basic stuff you try to teach freshmen. And, oh yeah, everything created excess heat.”

34

THE

PRESS

“The world does not yet know conclusively whether Stanley Pons and Martin Fleischmann will enter science’s pantheon with Ernest Rutherford, one of the giants of 20th-century physics,” read The Wall Street Journal editorial on Tuesday morning, April 12. ‘““The world, however, is getting a rare look at the clash between science and the politics of science—a clash between human curiosity’s innate optimism and the compulsive naysaying of the current national mood.” Ron Parker of MIT, among others, would later say that he heard in this editorial vaguely insulting echoes of Spiro Agnew’s immortal ‘‘nattering nabobs of negativity.”’ In any case, the Journal then redefined this clash as the “‘creative impulse of a Fleischmann and Pons” contending with “what might be called the ‘Academy Mentality.’ ”’ To the Journal it had become a battle of good against evil, of progress against stagnation, and, thank goodness, good had triumphed. The A&M and Georgia Tech confirmations had clinched it: “Clearly, scientists have once again discovered something new under the sun.” With the help of The Wall Street Journal, cold fusion began to play like a nuclear version of the emperor’s new clothes. On one side were those scientists who believed, as the Nobel laureate Luis Alvarez once put it, ‘Only trust what you can prove.” These scientists pointed out whenever possible that nobody had shown any data to support cold fusion, let alone prove it, and that certain basic experimental procedures had been consistently ignored in the pursuit of salvation. These scientists, however, soon

found themselves firmly entrenched on the wrong side of the paradigm: if they did not embrace cold fusion, it was because they were hopelessly self-interested. In the most egregious example of this defense, Dana

198 #9 GARY

TAUBES

Rohrabacher, a California congressman, later wrote a Los Angeles Times editorial in which he labeled the anti—cold fusion faction, which could

be called the pro-science faction, “‘small, petty people without vision or curiosity.” There was, of course, something of a catch-22 inthis attitude: if you knew enough nuclear physics to understand why cold fusion appeared to be dead wrong, you were, by definition, sadly attached to the old paradigm. Thus small, petty, and lacking in vision. If you knew little, nothing, or absolutely nothing about nuclear fusion—as did, perhaps, the distinguished representative from California—then you were considered a visionary and might get funding from the Golden State, if not the federal government, to pursue research in the subject. It was Thomas Kuhn of MIT, the philosopher of science, who first elucidated the concept of paradigm shifts in The Structure of Scientific Revolutions. On accepting new paradigms in science, he wrote: How, then, are scientists brought to make this transposition? Part of the

answer is that they are very often not. Copernicanism made few converts for almost a century after Copernicus’ death. Newton’s work was not generally accepted, particularly on the Continent, for more than half a century after the Principia appeared. Priestley never accepted the oxygen theory, nor Lord Kelvin the electromagnetic theory, and so on.

Kuhn then went on to catalog the great scientists who themselves were aware of what he called the ‘‘difficulties of conversion.” Charles Darwin wrote at the end of his On the Origin of Species: Although I am fully convinced of the truth of the views given in this volume ..., I by no means expect to convince experienced naturalists whose minds are stocked with a multitude of facts all viewed, during a

long course of years, from a point of view directly opposite to mine. . . . But I look with confidence to the future,—to young and rising naturalists, who will be able to view both sides of the question with

impartiality.

Max Planck later wrote what many believe are the definitive words on the subject: A new scientific truth does not triumph by convincing its opponents and making them see the light, but rather because its opponents eventually die, and a new generation grows up that is familiar with it.

BAD

SCIENCE

8 199

From April 12 on, any physicist who criticized cold fusion was considered among those whose death was—with sincerest regrets—inevitable, after which the new cold fusion paradigm could flourish. As Pons later told Time magazine: “‘Chemists are supposed to discover new chemicals. The physicists don’t like it when they discover new physicals.”>¢ This battle over the quality and source of the science in cold fusion led to an outbreak of anti-intellectualism. The physicists were on the wrong side of the paradigm not only because they understood nuclear fusion but also because they had established themselves over the years, in the words of New Scientists pseudonymous columnist, Ariadne, as clever dicks. Ariadne discerned a “sneaky hope” in cold fusion and the frenzy that followed: There is a lot of wicked appeal in the idea of decades of work and untold sums of money spent on complicated apparatus trying to capture the violence of the Sun’s interior being proved entirely unnecessary, indefensible though the attraction is. I suppose it is the primitive satisfaction of seeing a bunch of clever dicks being shown not as clever as all that and unable to see what was under their noses all the time. There are few such moments to relish in the daily round or the common task, let alone in such numinous enterprises as the pursuit of fusion. And so it went. Burton Richter, who shared the 1976 Nobel Prize in

physics and was as smart as they came, later described the public response this way: “These snotty physicists, what the hell do they know? Stick it in their ear. We showed them.” Newsweek later referred to the physicists as the “‘self-anointed high priests of science.”’ In the same article, the victorious chemists were the ‘mere beaker keepers more suited to plumbing the wonders of polyester than the mysteries of the universe.”’ In the hierarchy of science, physicists had risen to the priesthood during World War II with the Manhattan Project. This had procured them funding beyond the dreams of their brethren and labeled them forever as the makers of the atom bomb, what Richard Rhodes would

Miltonically call ‘‘death into this world and all our woe.” Indeed, a half century later roughly half of all physicists, whether theorists or experimentalists, were supported by the Department of Energy, which also supported the nuclear weapons programs. Why this is so is a little un-

clear. Dennis Flanagan, a former editor of Scientific American, has suggested that “the public and its elected representatives have the understandable impression that physicists, having once created a revolutionary weapon, may do so again. A country has to be nice to physicists or maybe somebody else’s physicists will get the drop on you.”

200 & GARY

TAUBES

Now no one had to be nice to the physicists, because the chemists had

gotten the drop on them. That chemists had perhaps the worst public image of any scientists—what with toxic waste dumps and carcinogens being perceived unfairly as their major contribution to humanity—was simply indicative of the power of this primitive satisfaction. These chemists were elevated by the grace of Pons and Fleischmann to the exalted position of underdog. Not that all chemists were inherently credulous, by any means, on the subject of cold fusion. But, after April 12, any chemist who saw reason to criticize cold fusion was considered by its proponents to be, ipso facto, a closet physicist. And that was that.

35

ERICE,

SICILY,

AND

DALLAS,

TEXAS

The timing of The Wall Street Journal pro—cold fusion editorial on April 12 had been as curious as the message was astonishing. That same day two major cold fusion meetings took place—a world conference on cold fusion in Erice, Sicily, and the Dallas American Chemical Society meet-

ing; between them Jones, Fleischmann, and Pons would all present data. Rather than wait to see what came out of Erice and Dallas, the Journal editors opted to declare cold fusion confirmed a day early, thus establishing their position as a participant in the blossoming science and not just a reporter of the facts. The ‘“‘world conference’’ was the brainchild of Antonino Zichichi, a

Sicilian physicist cum media star whom the Italians viewed as a dapper version of Carl Sagan. Zichichi ran a series of one-week academic programs in Erice that were popular with the academic community, if for no other reason than the spectacular setting. When cold fusion broke, Zichichi decided to give the European community its first arena in which the two paradigms—hot fusion and cold fusion—could fight it out. Zichichi seemed to favor the new paradigm. He seemed to believe that the European community should dump its extraordinarily expensive hot fusion plans and begin building cold fusion reactors. ““Europe, the Soviet Union and the United States are about to launch a ten-year, $1,000 million project to study the viability of fusion,” wrote Nature, “but all may change pending the conclusions of the conference.”’

BAD

SCIENCE

# 201

The location was Zichichi’s Ettore Majorana Center for Scientific Culture in Erice. Steve Jones had been invited by T. D. Lee, a Nobel Prize winner at Columbia, who told him that Martin Fleischmann would

be coming and they’d need Jones for balance. Kelvin Lynn, whose Brookhaven experiment had been reported prematurely by the press as a confirmation, had been called directly by an Italian scientist who said he was the head of Italy’s atomic energy commission and had read about Lynn in The Wall Street Journal. Zichichi also invited Dick Garwin of IBM. Garwin suggested that they would be better served by waiting a month, but Zichichi disagreed. Zichichi also recruited several older Soviet physicists, including S. S. Gerschstein, who had done the original work on calculating the cold fusion rate, and Leonid Ponamarev, the

expert on muon-catalyzed fusion. Many of the participants heard about the Texas A&M and Georgia Tech confirmations only on the day they left for Italy. Thus the general mood, even among the physicists—like Steve Koonin, who would present his theoretical calculation, and Dick Garwin—was that maybe something was happening in these sophisticated test tubes that they simply did not understand. When they arrived in Italy, they were greeted by pictures of Jones and Fleischmann on the front pages of the newspapers under enormous headlines: FUSIONE FREDDA. On the morning of the twelfth, there were rumors coming out of

Frascati that Italian scientists had duplicated cold fusion as well. “‘So the Italians had done it,’’ Koonin would say, ‘‘and everybody was happy.” Rumors also circulated that physicists in Moscow had seen neutrons; TASS, the Soviet news agency, was reporting that Professor Runar Kuzmin and colleagues at Moscow University ran twenty experiments and “confirmed the experiments of U.S. colleagues to obtain nuclear fusion at room temperature.’ Kuzmin et al., at least in the West, would

not be heard from again on the subject. The Sicilians themselves seemed to be greeting cold fusion as the energy source of the future. The governor of Sicily made an appearance at the meeting and announced that he was prepared to turn an old American military base into a center for Italian cold fusion research. At 8:30 in the morning, the courtyard at Zichichi’s institute was already filled with reporters, TV crews, and photographers. Fleischmann then gave his standard presentation, similar to his CERN seminar, and was unable to answer the key questions. Koonin made a point of asking about light water controls. Fleischmann responded that he was not pre-

202 B GARY

TAUBES

pared to discuss it. Isn’t it the essence of good scientific procedure to run controls? Koonin was thinking. What the hell is going on? Jones also gave his standard presentation. Garwin reviewed all the possible and all the definite sources of error in cold fusion experiments. He suggested that neither Fleischmann nor Jones could rule out what he called the “‘arcs and sparks” hypothesis as the source of their neutrons. As he later explained, the hypothesis “‘says that when you are looking for thermonuclear reactions . . . very often, instead of heating up the entire bulk of hydrogen, the energy goes into heating a very few particles to very high energy. If I have a spark and I have ten kilovolts on a few of the particles, I will get plenty of neutrons. I can’t make energy from that because it is well known that the energy loss to the other cold particles is much bigger than the energy generated.” Garwin wanted to know why there wasn’t a vast amount of data since March 23, 1989, confirming the BYU

observation of neutrons. This is

one of the unwritten laws of experimental physics: if a signal is real, it will get larger with time and effort. Although there had been sporadic reports of confirmations of the Jones effect, there had not been the avalanche of data one would have expected. What mystified Garwin was the confidence with which Fleischmann presented his data, considering the criticisms he was hearing. And the BYU experimental data were not nearly so good as Jones insisted, yet Jones seemed serenely confident. Garwin had agreed to write a commentary for Nature on the Erice meeting, but all he could report was that for the results to be right would demand a “‘multi-dimensional revolution”’ and he would bet against that. He confessed to his daughter, Laura Garwin, that he found it all very confusing. ““The idea of wanting fame or overinterpreting one’s result is quite foreign to him,”’ she said later. “He said on the evidence they had presented, he didn’t think they had anything. Yet they were so confident. He believed that they must be holding something back. That was the only explanation he could come up with. Because if he had that data he wouldn’t be so confident.” The Dallas meeting was no more a success than Enice, but at that point in the cold fusion affair, simply the presence of Pons or Fleischmann was a cause for celebration. When Stan Pons arrived in Dallas, he received

the kind of treatment usually reserved for rock stars. The Deseret News reported that Pons had checked into town under an assumed name and was never seen around the conference without an enormous bodyguard. The exalted atmosphere had a precedent in the spring 1987 meeting of the American Physical Society, during which 3,500 physicists tried to

BAD

SCIENCE

§ 203

elbow their way into the ballroom of the New York Hilton—2,000 succeeded—to hear about the breakthrough in high-temperature superconductors. This meeting became known as the Woodstock of physics. Once cold fusion broke, the chemistry community hoped that cold fusion would do for them what high-temperature superconductors had done for condensed matter physicists. There were, however, a few differences between the two meetings, chief among them being that the physicists who had crammed into the New York Hilton had been discussing a phenomenon they knew to be real, while the chemists in Dallas persistently talked around one they hoped was real. In fact, the American Chemical Society had a bylaw prohibiting the presentation at a technical meeting of results that had not been peerreviewed in some form. The purpose of the rule was to discourage the dissemination of specious results. However, the ACS executives felt that the publicity had made cold fusion a special case. Clayton Callis, the ACS president, said he had begun to arrange the Dallas session on March 25, the day after he read about cold fusion in The Wall Street Journal. The cold fusion session was scheduled for lunch on the twelfth, and

it was moved to successively larger auditoriums until the ACS finally settled on the basketball arena at the Dallas Convention Center, which held 10,000. The conventioneers estimated the cold fusion crowd at

7,000. No one was allowed to photograph or tape-record the proceedings, although the ACS did videotape them. (Copies of this ‘‘historic session,” as 1t was advertised, were sold for $400 each.) The press was

seated in an upper balcony, where they’d be too far away to disturb the session by asking unscientific questions.*” The tone, which could be described as self-congratulatory rather than

scientific, was set by Valery Kuck, the session chairman, who began by announcing that she had just been informed that Moscow had confirmed. Then Callis took over and explained that billions of dollars had been spent on hot fusion programs. “Much has been learned about plasma physics,” he reported, ‘‘and, while much progress has been made, the goal has remained elusive and the large, complicated machines that are involved appear to be too expensive and too inefficient to lead to practical power.”’ Callis suggested that ““chemists may have come to the rescue,’ which provoked a burst of applause and some laughter. Callis had had little trouble enticing Pons to speak, as Pons had been anxious to present his data to an audience of his peers. The major problem confronted by the ACS was that it couldn’t count on many of the chemists understanding the nuances of fusion or even, for that matter,

electrochemistry. To prepare the audience for Pons’s lecture, Al Bard

204 § GARY

TAUBES

was brought in to explain the electrochemistry with Ernest Yeager, a distinguished chemist from Case Western Reserve University, and Har-

old Furth, head of the Princeton plasma fusion program, was asked to discuss nuclear fusion in general. Furth would be problematic because he understood. fusion; he had been working on it for over thirty years and had been director of the Princeton hot fusion program since 1981. Although he wouldn’t admit it, publicly he seemed disturbed by the fanatic goings-on. Furth strongly suspected that cold fusion was a “nutty thing,” as he would say, but he had learned to keep his suspicions to himself. ‘“Even if one is ultimately right,” he said, “‘it’s an awkward position to be in. It suggests a lack of open-mindedness. So one shouldn’t rush to judgment, at least not publicly.”” His public technique was to make what he called tangential contributions. He would simply suggest experiments that would help “the perpetrators clear the air.”” The night before the conference, Furth had a phone conversation with Nate Lewis, who briefed him on the lack of light water controls and some of the other problems. Now, after explaining the physics of conventional fusion, Furth suggested several times that Pons consider doing control experiments with light water in order to be taken seriously by the scientific community. Pons and most of the chemists present dismissed the suggestion. Perhaps they considered this a technicality. Furth also lectured on the nuclear radiation one would expect from fusion. He tried to communicate the troublesome fact that even if one could think of ways to produce fusion without neutrons, they wouldn’t explain why the more favored fusion mechanisms, which always produced neutrons, would be so mysteriously suppressed. When asked whether anyone had a viable theory for this, Furth answered, in short, ‘“‘No.’”’ He then suggested again that theoretical nuclear physicists would not try very hard to find such a theory “until the American Chemical Society repeats this experiment and shows that while it works in heavy water it does not work in light water. The moment you do that you’ll have the best theoretical nuclear physicists in the country finding solutions.” Furth was ignored again. In fact, the Dallas meeting and Furth’s political tightrope act added to the growing conviction among some proponents of cold fusion that the hot fusion community was simply too self-interested to face the dawn of a new paradigm. That Furth simply had more faith in physics than he did in Stanley Pons seemed like just the kind of coldhearted rationale one would expect from a physicist. Pons then gave the same talk he’d given at Utah and at the University

BAD

SCIENCE

§ 205

of Indiana on April 4, with the same data, or lack thereof. What made

his Dallas talk notable was that it was received in an atmosphere of blind faith that would pervade cold fusion. In the question-and-answer period that followed the talks, several chemists asked pointed questions, and Pons’s uninformative answers were accepted with polite nods. When an MIT chemist, for instance, asked if he had tried any control

experiments with light water, Pons responded with abrief ‘‘Yes.’”’ Then, after a lengthy pause, he continued: “Several people are looking at this right now, including ourselves. . . . The confinement parameter which we have . . . that sort of reaction might be interesting.”’ This certainly seems to have been a meaningless response, yet it was not pursued. They simply moved to the next question. Then alittle later a chemist from Berkeley suggested that ‘“‘the energy resolution [of the gamma ray spectrum] appears quite good, actually better than expected with sodium iodide detectors.” This question was tantamount to an accusation of fraud: in science, data that are too good are a likely sign of fraud or error. Pons ignored it. He showed no righteous indignation. It’s conceivable that he didn’t understand it. Valery Kuck simply proceeded to the next questioner. She may not have understood it either. Carlos Melendres, a chemist from the Argonne National Laboratory, then explained that he and his colleagues had recalculated the table in the Pons paper that listed the prodigious numbers for excess heat as a percentage of break even. “Despite the lack of information about how the calculations were done,” Melendres said, he and his colleagues were able

to calculate what the numbers should have been, and they were all different from what Pons and Fleischmann had reported. In fact, he said,

the excess heat seemed to fall always within the range of what might be expected for heat released from the recombination of deuterium and oxygen in the cell. “So we don’t find really anything exciting about it,” Melendres said. ‘‘Maybe we’re missing something and can compare notes with your calculations.” “We'll take a look at them,” Pons replied. “‘[But] I don’t see how you can obtain fifty megajoules of energy.” Melendres suggested that they go through the numbers, choosing a particular example from Pons’s paper, in which, Melendres explained, the value of excess heat was only a quarter of the break-even point. It was actually less than the heat that they would expect to get from recombina~ tion. “So we don’t see .. .”’ Melendres said. Pons was obviously confused. Melendres repeated the numbers.

206 & GARY

TAUBES

“Okay,” Pons said finally, “‘I’ll be happy to look at that. [I] can’t do the calculation right now.” No one pressed Pons for a legitimate answer, nor did anyone seek out Melendres afterward to follow up the point.** Kuck, meanwhile, moved

to the next question. For Harold Furth, the Dallas meeting was a surreal experience. He later observed that he felt as if he were being subjected to the kind of experiments devised by the psychologist Solomon Asch. Asch would take one genuine experimental subject and seat him with six confederates, all of whom

had been instructed in advance

to answer

simple

questions incorrectly. Asch would then show all seven a set of straight lines, of which one would be distinctly longer than the rest. To the question which was the longer line, the six confederates would unanimously choose the wrong one. “It’s an amazing phenomenon,” Furth said. ‘‘The odd person, the true experimental subject, begins to think maybe he’s wrong after all, even though he knows perfectly well which line is longer.”” Asch found that over a series of experiments, three out of every four genuine subjects chose to side with the group. In any case, the press treated Dallas as a triumph for Pons, reporting that he received a standing ovation from the 7,000 in attendance. Even the skeptical New York Times seemed impressed by the enthusiasm and predicted that cold fusion had ‘“‘opened up a schism between chemists and physicists that may take years of experimentation to resolve.”

36

DALLAS.

COLLEGE

STATION

When Chuck Martin went off to the Dallas meeting, he left his cold fusion work in the hands of his researchers, Frank Cheng, Del Lawson, Mike Tierney, and John McBride, and asked them, as Lawson put it, “‘to

figure out what the hell was wrong.” All they knew for certain was that they had a cold fusion cell that would generate excess heat regardless of what kind of electrode or electrolyte was used. Cheng had been of the opinion all along that Martin, Marsh, and Gammon, in their pursuit of glory, had kept the younger researchers away from the experiments, allowing them, at most, to baby-sit. There was even some question among these chemists whether their names would be on any publication.

BAD

SCIENCE

# 207

Lawson felt they were used only as “human computers” to take down the data. And whether it was true or not, Martin, Marsh, and Gammon

managed to convey their impression that they doubted the experimental skills of these younger men, which made their work a matter of pride as much as anxiety. “When they left for Dallas,”’ Cheng said, “‘they left the cell to us, and

we did the sort of things you would normally do to troubleshoot an experiment, observations, assumptions you would make and test out.”

Among Cheng’s suggestions was to turn off an automatic stirrer they had used in the fusion cell to keep everything at the same temperature. It also took the oxygen and deuterium bubbles evolving off the electrodes during electrolysis and scattered them around the cell. With all the effervescence, it looked like a fresh glass of soda water. When Cheng turned off the stirrer, they could see that these bubbles had also been forming on an electronic quartz thermometer that was in the cell to double-check the temperature. In other words, the thermometer was somehow serving as a second cathode. This meant that they were putting more power into the cell than they thought. Unplug the thermometer, and the excess power went from 80 percent, which was the “‘striking”’ result they had announced publicly, to zero. They told Martin immediately. He called in from Dallas and at first refused to believe it. When they told Bruce Gammon, his first comment was “Damn.” Still, as Cheng said, ‘‘At that point all of us thought Pons and Fleischmann’s results were legitimate, and all we had to do was go back and reproduce the experiment and that would cover everybody’s ass.” Indeed, there had been one experiment, one cell, that had been generating

over 110 percent excess heat. Subtract the 80 percent for the thermometer, and that still left them with 30 percent excess heat to explain away. Martin held on to this thought for the next two months. He later recalled Uzi Landau, a Case Western Reserve chemist who would also report excess heat, saying to him, ““You have a pen in your hand. If one day the pen becomes a gun and shoots someone, you couldn’t say, Well, it’s been a pen for the ten years you’ve owned it.’’ So Martin still had one set of data which seemed to indicate excess power generation even on top of that added by the thermometer. He could not prove that all his results were compromised, only that all but one were. Still, Martin’s story is a stark example of the danger of speed in science. Within twelve hours of his press conference, he knew his experimental

results were compromised. Martin also came to see it as an unforgettable lesson about science and the press. ‘““Talking to the press is wrong,” said

208 @ GARY

TAUBES

Martin, ‘‘very wrong. It’s too easy. And the press can’t filter. They can’t tell whether the thing I’ve said is bullshit or right.”” (On Apmil 10, 1990, exactly one year after his press conference, Martin awoke feeling, as he put it, “‘like someone was sticking hot pokers down my throat.”” Sometime later that day he realized that this was the first anniversary of his announcement.

Thus, he said, ‘‘It must have been my subconscious

making me pay for my stupidity.’’)

37

ATLANTA

It started as just another one ofthose phone calls: how come the Georgia Tech Research Institute got neutrons and we didn’t? On April 12 a physicist at the Lawrence Livermore Laboratory who was trying to replicate cold fusion spoke to Gary Beebe, who was a research scientist at GTRI working with James Mahaffey. Livermore and GTRI had virtually identical experiments, down to the boron blocks and the borontrifluoride (BF;) neutron detectors. The Livermore physicist told Beebe that they had observed anomalous readings of the kind GTRI had taken public, and they had also thought to take the counters out and put them behind the boron blocks. There was one fundamental difference, however. It seems that when

Mahaffey had taken the detector out, he had handled it very carefully, daintily even, by the cord. The Livermore scientists may have been less concerned about the health of their detector, because they had simply picked it up in their hands. One of them noticed that when he did that, the count rate shot up as though he had walked into a shower of neutrons. Either cold fusion depended on someone holding a neutron counter in his hand, which seemed off the wall even for cold fusion, or

the detector was heat sensitive. The guy from Livermore told Beebe that they took a heat gun and zapped the BF; counters and discovered that they were hypersensitive to heat. In fact, these counters were so sensitive to heat and humidity that they could almost be used to measure the temperature of a room. This seemed particularly true of the old BF,’s that GTRI had obtained. The heat and humidity sensitivity of the BF, detectors was well known among physicists who used these detectors regularly. Regrettably, nobody had both-

BAD

SCIENCE

&@ 209

ered to inform the newcomers to this art who suddenly needed to count neutrons for cold fusion experiments.?? Beebe listened to the report from Livermore with mounting horror. Then Beebe, Rick Steenblik, and Darrell Acree repeated the Livermore experiment with their own counters. They held them in their hands; the count went up. They put a beaker of hot water near a detector; the count went up. Acree called Mahaffey. ‘‘Jim,’’ he said, ‘‘we’re all dead men.” Mahaffey observed later that Acree ‘‘wasn’t far from wrong.” And he added, ‘““We’d made this huge splash across the world that we’d found neutrons, and we hadn’t found neutrons at all.”

Mahaffey and his each day. One of conversation, again you got to look at

colleagues were still fielding hundreds of phone calls them was from Stan Pons. Mahaffey recalled the evoking a strong sense of déja vu: “‘We said, ‘Pons, your neutron detector. These things are extremely

temperature sensitive. At this low neutron rate, what’s normally noise

gets counted as if it’s neutrons. You have to be careful about the temperature dependence.’ So we asked himif he had been careful. He said, ‘No, I don’t have to.’ It was really strange. It was like he was on something. He said, ‘No, I don’t have to do that.’ We thought, Why doesn’t he have to? Anyway, he never mentioned that neutron detector ever again.”

38

CAMBRIDGE

On April 12, while the Woodstock of cold fusion was beginning in Dallas, the public relations office at MIT released a short statement: Peter Hagelstein, an MIT electrical engineer, had concocted a theory to explain cold fusion; he had applied for patents “‘in connection with this theoretical analysis. The News Office cannot discuss the specific technology prior to the publication of the journal articles,” the statement went on to say. “Further details will be provided upon acceptance of the papers.” It was all pretty mysterious, and Hagelstein did not make it any less so. He was well known among the scientific and defense communities, and so was his life story. The MIT press release noted that Hagelstein had been one of the key figures in the development of the nuclear-pumped X-ray laser for the Star Wars missile defense program. He had pioneered

210 © GARY

TAUBES

this top-secret technology as a graduate student a decade earlier. Hagelstein had a distaste for publicity of any kind and shunned the media on principle. Perhaps for this reason reporters referred to him in weighty but similar terms: William Broad of The New York Times, who had helped deify Hagelstein in his book Star Warriors, called him “‘an elusive brooding genius.” Philip Hilts, then of The Washington Post, called him “a reclusive scientist widely regarded by colleagues as unusually brilliant.” And so on. Physicists referred to Hagelstein’s history as the Hagelstein myth and talked about it as a cross between Dr. Strangelove and Faust. One Star Wars researcher described it as ‘‘a myth of scientific creativity warring in one person against morality . . . the image of Dr. Strangelove with the hand coming up to his neck and the other hand trying very hard to push it away.” Hagelstein had earned both his undergraduate and graduate degrees in electrical engineering at MIT. He then went to the Lawrence Livermore Laboratory, lured by a sizable scholarship and the charm of Lowell Wood, who was considered by the more liberal scientific community to be the evil genius, along with Edward Teller, of the Strategic Defense

Initiative. At Livermore, Hagelstein developed a code for the analysis of X-ray lasers, and, almost against his will, so the myth had it, started thinking about problems of nuclear-pumped X-ray lasers. In a stroke of insight—again, almost in spite of himself—he came upon the idea that would theoretically make them work. The myth, like all good myths, also had a love story attached to it. Hagelstein was deeply involved with a woman of strong moral principles, who broke off the relationship because he wouldn’t renounce his obsession with his immoral weapons research. She took to picketing Star Wars on the MIT campus. Hagelstein listened to requiems that echoed his melancholy. That this aspect of his life had become public knowledge may have explained why the brooding, elusive genius declined interviews. Finally Hagelstein broke away from the pull of Livermore and was offered an MIT associate professorship, which was considered a high entry-level position for an academic. Hagelstein had apparently been visiting Livermore when cold fusion broke. Edward Teller, the director emeritus of the laboratory, was optimuistic about cold fusion. According to George Chapline, a veteran Livermore physicist, Teller passed this optimism to Lowell Wood, who conveyed it to Hagelstein. Hagelstein then started working on his theory, using the same theoretical methods that had been successful for the X-ray laser. Meanwhile, Wood’s group began cold fusion experiments.

BAD

SCIENCE

#@ 211

The first weeks after the Utah announcement were full of rumors of cold fusion—induced explosions at Livermore. The computer network reported that “‘they had a small accident when their experiment blew up. It destroyed the fume hood it was in and blew a four inch hole in the concrete beneath it.” It sounded suspiciously like Pons’s seminal meltdown. Later the San Jose Mercury News reported that the Livermore experiment blew up with a loud pop, citing unidentified sources at the laboratory. The lab then officially announced that the rumors of an explosion were “‘exaggerated,”’ which did not mean it hadn’t happened. Will Happer, a Princeton physicist who was also head of JASON at the time, later explained that, like a lot of other labs unfamiliar with electrol-

ysis, Wood’s group had a small hydrogen explosion. ‘These things are a bit dangerous,”’ Happer said. “You don’t recombine the hydrogen and the oxygen; it can explode—fracture the glass and stuff.” In any case, when Hagelstein went public, he seemed to have more faith in the Livermore optimists than in the MIT pragmatists. Ron Parker said Hagelstein never bothered to look at the experiments that were being run at the Plasma Physics Center, nor did he apparently talk to any of the physicists or chemists involved. Instead, Hagelstein held a seminar with six of his colleagues. John Deutch, the provost at MIT and a former director of energy research at the Department of Energy, was one of them. Deutch later said he thought Hagelstein’s theory was “‘malarkey,” but apparently he didn’t bother to say so at the seminar. When Hagelstein announced that he wanted apress release and patents, Deutch went along because it was ‘‘absolutely essential to a university not to reach a single judgment about what could be true or not true.”’ That one of his faculty was going to pursue a possible explanation for cold fusion struck Deutch as “being the essence of the university.’ Deutch was quoted in the press release saying, ‘“We are pleased to see Professor Hagelstein proposing an explanation for ‘cold fusion,’ and we are encouraging investigators both here and at other research institutions to continue their work on this most surprising phenomenon, which may have enormous consequences.” This was a minor work of academic double-talk, but it seemed to give an entirely different cast to MIT’s official position, which was that it had no official position. Science, after

all, is not determined by administrative decree, but by working scientists. What Deutch hadn’t realized was that cold fusion had affected the political environment. It was like a gas leak in an otherwise comfortable home. The simple act of lighting a match comes to have far-reaching consequences.

212 © GARY

TAUBES 39

LAFAYETTE,

CALIFORNIA.

WASHINGTON,

AN

ENERGIZED

QUEST

SKEPTICISM

FOR

NUCLEAR

IN RUSH

D.C.

FUSION:

OF EXPERIMENTS,

OPTIMISM’ REPLACES THEORIES

Front-page headline in 7he Washington Post, April 13, 1989

On the morning of April 13, officials at the White House called Glenn Seaborg. The UC Berkeley chemist had won the Nobel Prize in 1951 for his work on transuranium elements, which included the discovery of plutonium. Seaborg was eating breakfast in a restaurant in Lafayette, California, and the fact that the White House tracked him to his local

eatery was some indication of the urgency of the call. Seaborg was asked to catch a noon flight to Washington and brief President Bush and Chief of Staff John Sununu on the possibilities of cold fusion. Seaborg had known Bush since the Nixon administration, when he had been head of

the Atomic Energy Commission (now the Department of Energy) and Bush had been ambassador to the United Nations. Seaborg arrived in Washington that night and was met at the airport by Robert Hunter, the DOE director of energy research. Hunter was trying to assemble a summary of the cold fusion situation. His office had just called all the DOE laboratories asking for progress reports. Ron Parker of MIT got a call at six o’clock that evening from Hunter, who said they needed a cold fusion report that night. Parker’s crew went to work and faxed the report to him at one o’clock in the morning. On the morning of April 14, Seaborg met first with James Watkins, the secretary of energy, then with Sununu. Both men, along with other congressional leaders and all members of the Senate Energy and Natural Resources Committee, had been formally invited by the Utah congressional delegation to fly to Salt Lake City on Apmil 29 to witness a private cold fusion demonstration. Finally Watkins, Sununu, and Seaborg entered the Oval Office to brief

President Bush on fusion in a bottle. Whether the president and his staff were optimistic that cold fusion was the answer to America’s energy problems is no longer clear. Certainly Seaborg was not, which was the message of his briefing. He told the president that it was “‘a very small effect if any effect at all.” Nonetheless, Seaborg did suggest that the DOE assemble a panel of scientists to examine the cold fusion claims officially.

BAD

SCIENCE

§ 213

Seaborg did not believe there was anything to the claims, only that it wouldn’t be adequate, considering the publicity, “just to deprecate the work and say they didn’t believe it.”

40

SALTLAKE

CITY

The budding cold fusion interests in Utah took the Hagelstein-MIT announcement as tantamount to a declaration of war. Chase Peterson told reporters that it was time to play “hardball.”” And Paul Van Dam, the state’s attorney general, announced that $500,000 of the $5 million

the state had allocated for cold fusion research would go to patent attorneys, and his office, as the university’s legal representative, fully intended to spend every last penny. Somehow this was not difficult to believe. Van Dam said that they were prepared to protect cold fusion for the state of Utah against all comers, and that they would ‘‘do whatever it takes; On April 13, one day after Hagelstein went public, Stan Pons announced that he was rejecting the offer of support from Ryszard Gajewski’s office at DOE. Instead, he would stay with his friends at the Office of Naval Research. No reason was offered by either Pons or the university. “It is a scientist’s prerogative,” said Pam Fogle of the Utah public relations office. A DOE spokesperson, when informed of the rejection by a reporter, replied simply, ““You’re kidding.”’ Pons would later testify in Congress that he rejected the DOE offer because he and Fleischmann had already managed to do much of the proposed work during the “considerable delays” in the review process. Thus, he said, “it would not be right for us at that point to take the money to do work that had already been done.”’ This sounded noble but rang a bit hollow. The administrators at BYU later insisted that Pons had rejected the DOE offer so that his original DOE proposal would not become a public document. The contents of the proposal, they believed, would immediately exonerate Steve Jones of any act of piracy.” The more likely explanation was that Pons had grown to distrust Gajewski and was less than enamored with some of the scientists under his umbrella. That Gajewski would also be supporting Hagelstein—to develop an extreme ultraviolet laser, at a cost of $1.2 million over three years—may have explained the timing of the rejection.

214 8 GARY

TAUBES 41

THE

NETWORK

The hot story on the network on April 13 was that cold fusion was not new. It began with a note from a computer scientist’ named Dennis Erwin Thurlow:

A swedish engineer patented a method for producing helium from hydrogen with Palladium back in 1926... . | got a note from someone else about this, and they said the process had never been a commercial success because it generated too much heat. Can someone with access to Swedish patent files check this? If true, it would be a great joke (if you like irony). The engineer in question was John Tandberg, a former head of the Electrolux lab in Stockholm. On February 17, 1927, Tandberg applied for a Swedish patent entitled “A method for production of helium.” Tandberg was refining the research of two University of Berlin chemists, Fritz Paneth and Kurt Peters, who had published an article in 1926

claiming that they had electrolyzed hydrogen in palladium and detected helium. This convinced them that they had created helium. It was a kind of nonfusion fusion claim, because neither the neutron nor deuterium

was discovered until 1932. Only months after publishing their results, Paneth and Peters retracted, explaining that the minute amount of helium they had observed could be attributed to contamination from the air rather than fusion. The helium measurement was atricky one, and it was understandable how they could have erred in their excitement. Sixty-three years later, Stan Pons would make a similar mistake. The saga was continued in a Swedish technical journal, Ny Teknik, whose account was translated and typed onto the net by a Swedish computer scientist. It explained the fate of Tandberg’s patent application: “The patent application was returned with the words ‘read, but not understood,’ and on Nov 17 the same year the application was rejected with the motivation that ‘the description is not so complete that a competent person could use it to use the invention.’ ’’ Ny Teknik seemed to believe that Tandberg had missed out on the discovery of the century, done in by an ignorant patent bureau. When Pons heard the story, he agreed.*? The network’s second big story of the day was in its own way even more remarkable. Two graduate students from the University of Washington, using a modern version of Paneth and Peters’s technique, joined

BAD

SCIENCE

# 215

the growing list of those whose reputations would be marred by quick and dirty cold fusion experiments. Their names were Van Eden and Wei Liu, and they held a press conference in Seattle to announce their discovery and take their place in history. With a dozen reporters and four television crews in attendance, Eden and Liu said that they might have discovered tritium in the gases emitted by a variation on a PonsFleischmann cold fusion cell. Not having platinum, they had used a gold cathode. The press release, which appeared quickly on the network, reported that Eden and Liu had been running their cell for three and a half hours,

using an instrument known as a mass spectrometer to sample the gases bubbling out of the cell. In this effluent gas, the mass spectrometer detected the presence of molecules of mass 5 and mass 6, which is to say molecules that had some combination of five or six neutrons and protons. Eden and Liu let their cell run for ten hours, while seeing these same molecules the entire time, then replaced the heavy water with light water, at which point the signal for masses 5 and 6 dropped away to nothing. The two intrepid graduate students interpreted these results to

mean that they had formed molecules of deuterium and tritium (mass 5) and tritium and tritium (mass 6), and thus had partially confirmed Pons and Fleischmann. They did not see neutrons, but they noted that their neutron detector was particularly insensitive.** A University of Washington computer scientist who attended the press conference and sent the press release out on the net, remarked:

Eden handled the leading questions very well and was quite pointed about the press making d premature sensation about the fusion work. Eden repeatedly gave very cogent discussions on how science works, and how confusion and cross-fertilization are important to the process. Let’s hope some of the meta-lesson on the scientific process makes it into the media. The network followed up with a disclaimer from a physicist at the university, who requested that readers reserve judgment on the tritium claim. He then suggested that Eden and Liu could be seeing D; molecules—deuterium, deuterium, and deuterium—which would be mass 6, or D,H molecules—deuterium, deuterium, and hydrogen—which

would take care of mass 5. ““Most people here feel that this was an unfortunate case of jumping the gun,” he wrote. “Of course they may be correct.” Next came a message from another Washington physicist, who was considerably less forgiving. Its subject was “‘awful fusion exp. by grad students in solid-state at the U.W.” He explained that “‘mass spectrome-

216 B GARY

TAUBES

ters are plagued by weird ions” such as DDH + and DDD + . He added that if Eden and Liu had detected tritium, which is radioactive in its own

right, this radioactivity “should have made even that puny detector light up like a Christmas Tree.”

42

FUSION

‘‘COHERENT BY

PETER

L.

THEORY’’

HAGELSTEIN

The invisible network of facsimile machines began its universal distribution of Peter Hagelstein’s four papers on the afternoon of Friday,

April 14.* In order to explain cold fusion, Hagelstein had invoked what he called coherent effects, which were a natural outgrowth of his X-ray laser work. This was the obvious tool for Hagelstein to try out on cold fusion, considering his history. He was attempting to do what physicists call phenomenology, which is to take the body of available data and fit a theory to it. As an electrical engineer by training, he may have been straying into new territory. Hagelstein’s papers promised salvation with the same gusto that Pons, Fleischmann, and Peterson had on March 23. They also made no sense,

which is to say that the physics in them was less than masterful. Hagelstein began with the assumption that Fleischmann and Pons had demonstrated nuclear fusion at low temperatures and then tried to identify a mode of fusion that would produce only helium 4, which no one had yet looked for in cold fusion experiments. If cold fusion cells generated massive amounts of helium 4, that might explain why no lab had yet found either neutrons or tritium, which was what orthodox physics expected from D-D fusion. Unfortunately, Hagelstein had made basic errors in his quantum mechanics, which was noticeable regardless of which side of the paradigm one stood on. One MIT physicist remarked that the papers reminded him of one of his students in an undergraduate course, ““who couldn’t do physics to save her life’? but would list a seemingly random sampling of equations and numbers in her test books and eventually arrive at an answer that was inevitably and totally wrong. “Somehow,” the physicist said, ‘if you had your glasses off, it sort of had the gestalt of physics. And that’s what this

BAD

SCIENCE

8 217

is like. He’s sort of rambling on, and he writes these formulas, and they look like quantum mechanics, and he comes up with an answer, and it’s got units like you expect, and it’s just not physics. It doesn’t have the ligaments to tie it together.”’ The head of the MIT physics department, Bob Birgeneau, was apparently as dismayed by the quality of the papers as he was by the popular belief that Hagelstein was an MIT physicist. An editor’s note was appended to the press release explaining that Hagelstein was not a physicist but “an electrical engineer and [a] computer scientist.” On the fourteenth Hagelstein held an open seminar to present coherent fusion theory to his MIT colleagues. Hugo Rossi, the dean of the College of Science at Utah, was passing through Boston at the time and took the opportunity to sit in. By now Rossi for one had become optimistic. The cold fusion situation in Utah looked good. Pons had told Rossi that it would take three weeks for labs to start confirming cold fusion, and now, three weeks later, both Texas A&M and Georgia Tech

had checked in, although the latter would soon check out. Rossi had also spoken to Jack Simons, a distinguished chemist at Utah who had come up with a cold fusion theory. Rossi was curious about whether Hagelstein’s would be similar. So he went enthusiastically to Hagelstein’s seminar. “I saw how Hagelstein’s colleagues accepted, or failed to accept, his presentation,” he said later, “‘and I began to get an ever-deepening sense of how difficult it was all going to be. “Hagelstein started giving the historical background. First he gave five reasons why cold fusion was impossible. Then he said, ‘Now Pons and Fleischmann have this evidence. On the basis of that, there’s no reason

for anybody to try to devise an explanation for what’s happening. However, there are these other confirmations.’ And at that point, Ron Parker said, ‘What other confirmations?’ And Hagelstein mentioned a few of them, and Parker said, ‘That’s all rubbish. There are no confirmations;

nobody has confirmed this. All we have is just this announcement.’ ‘Well,’ Hagelstein said, “Well, I’m just giving the historical part of my talk.’ And Parker says, “Yeah, but you’re trying to explain why you want to make sense of this, and that won’t do, because there is no such thing

as confirmation.’ .. . “Then Hagelstein described his theory. Then there was quite animated discussion, and Ron Parker took part in it, talking just about how improbable it was. . . . The number of events that could possibly transpire in this way were off by an order of 10° or 10? from what was claimed. And that was pretty much it.’’*°

218 © GARY

Those

TAUBES

who

tried to make

sense

of Hagelstein’s theory, however,

ended up trying to make sense of Hagelstein. Ron Parker later observed that ‘‘genius” is “‘an overused word.” He then speculated that Hagelstein was a victim of success at an early age. “Science 1s a field,” Parker said,

““where all kinds of different qualities make people successful, and being smart is a kind of prerequisite, but everybody’s sort of smart. Some people are geniuses if you just measure their IQ, but that doesn’t necessarily make people creative, or give people the verve to discover or be curious. So it’s only a necessary condition, being smart. It doesn’t guarantee success.” Others took the Hagelstein papers as a reflection of the kind of science done in the Livermore weapons laboratories. The nuclear-pumped X-ray laser, for instance, which Hagelstein had received so much credit for, had lately become something of a fiasco itself, complete with attendant accusations of fraud. As Robert Park, who ran the Washington office of the American Physical Society, was fond of saying, “The X-ray laser? I keep asking what X-ray laser? They haven’t made enough Xrays to take a chest X ray.” In Steve Koonin’s opinion, Livermore had always had a reputation as strong on hype and weak on science. He added that he had always liked Hagelstein, but “‘I think the hype part of Peter’s background took over alittle bit too much. I think you saw some of the Livermore egos coming through: the way Peter says, ‘I’ve got the answer,’ and doesn’t show it to anybody else. That’s the basic theme: Pons and Fleischmann and then Hagelstein. You think you’re doing this great thing, which is sort of at variance with everything else you know about the world—how could you not want to talk to everybody else to get their opinion and feedback?’’*” In any case, Peter Hagelstein’s debut on the side of cold fusion may have had as much of an impact on the future of the affair as any other single incident. In the last of his four papers, Hagelstein had outlined his assessment of ‘mechanisms for coherent fusion.” The outline covered fourteen points. The last three seemed to be the most informative, in that they reflected

how quickly Hagelstein had churned out his opus, and perhaps the influence of the Livermore egos as well. Hagelstein claimed that his mechanism could result in “breaking, melting, boiling, exploding or vaporizing” of the electrodes, which seems to mean it could account for Pons and Fleischmann’s meltdown, and maybe even the Livermore explosions. It could lead to “‘probable efficient energy conversion and associated catastrophic effects,’ which intimated that it would be a nifty mechanism out of which to build a nuclear bomb, not to mention “the

BAD

SCIENCE

§ 219

development of new fusion driven laser sources.” Finally, Hagelstein suggested, “Direct coupling of the coherent fusion energy into electrical energy with some efficiency seems to be likely.” All of this sounds very optimistic. But if MIT had applied for patents on the applications of Hagelstein’s theory—of which, Pons observed laconically, “I didn’t know you could patent theories” —it could be construed that the MIT patent factory was trying to lock up the royalties on every imaginable future use of Pons and Fleischmann’s invention.

43

ATLANTA

On April 14, James Mahaffey and his colleagues at the Georgia Tech Research Institute officially retracted. As Mahaffey would put it, they held the “negative press conference to try and cancel out our positive conference.’’ Hundreds of reporters were present, and they had trouble finding a room large enough to hold them all. Mahaffey reported that they had not, as previously believed, observed neutrons from cold fusion cells. He gave the most lucid explanation he could of what had happened and how they had erred. Mahaffey found the experience unforgiving, “like going to a hanging, where I was the hangee.”’

44

THE

PRESS

Four days after the two University of Washington graduate students Eden and Liu had announced the discovery of tritium, the story made The Wall Street Journal: TwO GRADUATE STUDENTS ADD A PIECE TO PUZZLE OF UTAH “FUSION” EXPERIMENT. The Journal article ended with a paragraph on the GTRI retraction. It said only that Mahaffey and company “‘may have discovered problems with a laboratory instrument they used to detect neutrons coming from a cold-fusion experiment.’ This was a peculiar, albeit gracious, way to

220 © GARY

TAUBES

say that the GTRI physicists had admitted that their previous results had been worthless. All of this could be considered nit-picking or Journal bashing, except that by now even the scientific community was depending on The Wall Street Journal for its information. As Hugo Rossi later, observed, “The whole project was so motivated by business people and business interests that one positive citing in The Wall Street Journal was worth a hundred [Steve] Koonins, if you could find them.” An inherent bias was at work in cold fusion that influenced the course of events perhaps more than the Journal staff realized. Jerry Bishop, whose beat now was cold fusion, explained later that positive results were newsworthy, while negative results were not: “I started off neutral. My only concern was to keep on top of it. I became biased in a sense that every time someone reported a positive result, I had a story. So I liked to see the positive stuff. It became a bit boring just reporting the negative stuff.” Bishop added that his bias may have been accentuated by the fact that he had tapped into the chemistry network, almost by accident, rather than the physics network. “If you go back and look at the coverage,’’ he said, ‘‘in those first few days, almost everyone

was on the radiation,

where the neutrons were and stuff like that. But when I got in to talk to the chemists, they weren’t interested in that. They wanted the heat measurements.” A year later Bishop was awarded the American Institute of Physics science writing award. The judges’ decision was based on the fact that cold fusion had been the story of the year, and Bishop had certainly covered it more vigorously than any other reporter. Many of the physicists invited to the ceremony apparently believed that Bishop’s rosy bias, which could also be called gullibility, was not a preferred characteristic of good reporting, even in science, and boycotted in protest. “It was a very weird lunch,” said one television journalist who was present, “‘because nobody was there.” In its April 17 issue, Newsweek also cast its vote for the reality of cold fusion. The magazine considered the remaining important question to be whether cold fusion would be a practical energy source. The article enlisted Stephen Dean, president of Fusion Power Associates, and Gerald Kulcinski, a physicist at the University of Wisconsin, to give the upside and the downside. On the upside, deuterium cost only about ten cents a gallon, and a “half-ton pickup could carry enough deuterium to power a 1,000megawatt fusion reactor for a year.”” The downside was that “‘palladium

BAD

SCIENCE

&§ 221

costs $5 million a ton (and the price has soared since the fusion report);

a 1,000-megawatt plant would need 400 tons.” This works out to $2 billion just for the palladium, assuming the price did not continue to rise. Kulcinski also pointed out that the higher the temperature, the less efficiently palladium absorbs deuterium, and power plants would have to work at fairly high temperatures. “But it’s something we'd have to engineer around,” Kulcinski said. Newsweek swept these caveats aside with a rose-colored kicker: “The history of fusion is littered with breakthroughs that weren’t. But as Pons and Fleischmann found, sometimes the long shots do pay off.’

45

SALTLAKE

GROUND

CITY.

ZERO

On April 17, twenty-five days and an estimated 1,000 phone inquiries after the announcement,

Stan Pons was back where it had all started,

facing a roomful of reporters in the Henry B. Eyring Chemistry Building. Pons had decided that the best way to make himself accessible to reporters, while leaving himself time to do some research, was to hold weekly

press briefings. Joining Pons for his first briefing were Cheves Walling and Jack Simons, two of his most prestigious colleagues from the chemistry department. Walling and Simons had recently derived a theory attempting to modify nuclear physics in order to account for their colleague’s problematic data. Their theory was a lesson in the philosophic principle known as Occam’s razor. The principle set down by the fourteenthcentury English philospher William of Occam, is, ““What can be done with fewer is done with more in vain,” or, Nature prefers simplicity. If Pons and Fleischmann had induced simple deuterium fusion at room temperature—and if Occam’s razor was valid—then the fusion would have generated helium 3 and neutrons, and tritium and protons. Nobody had seen enough of any of these to justify the Utah heat observations. So now Walling and Simons, as Hagelstein had done, had waxed slightly less simple postulating mechanisms that led to only helium 4 and heat, thus obviating the need to detect the ten millionfold more common fusion by-products.*?

222 © GARY

TAUBES

Simons was a theoretical chemist who had been at Utah since 1971. Walling was the most distinguished member of the department. He was an organic chemist who had taught at Columbia University before coming to Utah, and a member of the National Academy of Sciences. Pons had given both Walling and Simons copies of his paper early on, and they had found it, in Walling’s words, “compelling and striking.” They worked together to create a theory to explain the data. Walling knew little about nuclear fusion. He was, he said, just ‘“‘an old organic

chemist.” But he believed it was his responsibility to help his colleague explain the data. Simons had hoped to come up with a theory tantalizing enough to induce the theoretical physicists to carry the ball. So he proposed a mechanism that might explain the large release of heat with the infinitesimal release of neutrons. Walling told Nature their theory was ‘qualitative,’’ which is often a scientific synonym for vague. By the time of the press conference, Simons and Walling had written a theoretical paper entitled “Two Innocent Chemists Look at Cold Fusion” and sent it off to the Journal of Physical Chemistry. The Walling-Simons theory proposed that somehow the fused deuterium nuclei created a helium 4 nucleus, which then decayed by ejecting what is known as a “‘virtual” photon, which in turn would give up its energy to the lattice by a process known as internal conversion. The result would be helium 4 and heat and nothing else. Pons told the reporters that he was greatly cheered by the theory and believed that the two innocent chemists had accurately predicted his reported fusion rates. This was not surprising considering they had created the theory to do just that. Walling later remarked that he thought Pons was greatly cheered just by having allies from his own department who were willing to stand up and take the heat with him. Simons also said that their theory was “‘a helluva lot better than others I’ve heard of, including MIT’s,” by which one gathers he meant Hagelstein’s. This may have been true, but not because the Two Innocent Chemists’ theory made any sense. Dick Garwin later said he recommended to the editors of the Journal of Physical Chemistry that it be published with major changes: “It was highly original, of current interest; however, it can’t explain the heat or the billionfold-too-small accompanying radiation.” Simons sent a copy to Steve Koonin for his comments, having based some of the work on ideas in Koonin and Mike Nauenberg’s paper. Koonin wrote back, ““You have a real problem. These are all the right questions to ask. I don’t have any answers, and neither do you.” Koonin later added that Walling and Simons represented a type of

BAD

SCIENCE

@ 223

scientific illiteracy epidemic throughout cold fusion. ‘“They’re working in a field,” he said, “in which they a priori don’t know very much. They wnite papers that are really manifest nonsense, or wrong by twenty orders of magnitude, and they think they’re doing something great.’’®° In any case, the theory speculated a mode of fusion that would result in helium 4 and heat. It predicted that cold fusion cells should be generating atrillion helium four atoms per second, which was in the ballpark to explain the Utah results. So now at the press briefing, Pons announced that he had indeed detected helium 4 in the effluent gases of his cells, just as the Walling-Simons theory predicted. “Huge peaks,” he said. Nature called the helium 4 announcement ‘“‘a qualitatively new piece of supporting evidence.” Apparently, back in January, Pons had used a mass spectrometer hoping to find tritium and had not bothered to notice what else it might have detected. Once he heard from Walling and Simons, he went back to the data, only to find two peaks at mass 4. One peak he interpreted as deuterium molecules and the other as helium 4. This then explained why they hadn’t seen any neutrons or tritium. “It’s because we are producing helium instead,’ Pons said. As Simons argued to the reporters, “How can

you get helium 4 without a nuclear reaction? Chemistry doesn’t make helium.” Pons said they had ‘‘no doubt whatsoever about the helium measurement.” Unfortunately, helium 4 is garden variety helium. It is everywhere. Nate Lewis had spoken with Walling before the press briefing and asked him whether they had checked for the presence of nitrogen and oxygen peaks in the gases, which would have been evidence that they simply had air contamination in the sample. Walling had said no. Lewis found it unbelievable that they had forgotten or ignored such a simple check. (A week after the press conference, Walling presented Pons’s helium data at a meeting at the National Academy of Sciences. His audience, who knew considerably more about mass spectroscopy than he did, also suggested that the helium was contamination from the atmosphere. Walling later agreed. But he added, “‘About this time we also realized it was more

likely that helium would be stuck at the electrodes, so the thing to do was to analyze the electrodes.’’) Meanwhile, Pons also told the reporters in Utah that one of his cells had sustained a fusion reaction for 800 hours, which works out to 33

days. twice added more.

He said that the output to what they had claimed at that he now had four cells According to The Deseret

input was approaching eight to one, their initial press conference.*! Pons running and was outfitting nineteen News, Pons said he knew of thirty

224 BE GARY

TAUBES

laboratories that had positive results but were not going public because they were too busy filing patents. The New York Times reported that Pons knew of sixty such labs.

46LOS

ALAMOS,

NEW

MEXICO

Immediately following the briefing, Pons flew to New Mexico, where he was scheduled to lecture at Los Alamos on April 18. This was the big cold fusion day at the laboratory. Pons had two seminars scheduled: a small session with the lab’s scientists who were involved in reproducing the experiment and then an open seminar in the main auditorium. It was also expected that while he was in town Pons would work out details of a cold fusion collaboration with the laboratory. The Los Alamos scientists treated Pons with discretion. “‘He was quite nervous going into the situation,” said Rulon Linford, who was coordinating the cold fusion effort at Los Alamos. “‘I think people sensed that and didn’t become cross or angry or attacking.” Linford had spoken to Pons on April 14 in preparation for the seminar. He had suggested that Pons bring the life history of one particular cell. Pons would hear this request several times over the course of cold fusion. Linford had seen Pons’s presentation at Utah and had read the paper. ‘““What people were really interested in,’ Linford explained, “‘was trying to pin down this claim that the total excess heat was larger than could be explained by a chemical process. That of course implies that you have a full-time history of the net energy input and heat output from the cell, and none of that data had ever been exposed.’ Such a history would show the temperature readings on the thermometers as they rose and fell, the voltage and current readings minute to minute, as well as any bursts of gamma rays or neutrons. It would be a total accounting of all the energy going into the cell and all the energy leaving the cell for the cell’s lifetime. How else could it be determined whether the cell was truly generating more energy than it was consuming? Even the hard-core physicists at Los Alamos were willing to accept that the excess heat could be caused by some kind of fusion, provided Pons could prove the excess heat was real. This history was the kind of information required to do that. “I realized,”’ said Linford, ‘‘that he had

BAD

SCIENCE

#§ 225

all sorts of data from all kinds of cells. So I asked him to pick one cell that was particularly well understood and bring that data along.” Pons replied that he had this data and was willing to bring it and present it, but he never did. As Linford recalled: “When I asked him in the smaller group, initially he just deflected the question and kept going. I told him that if he would present this data in the very beginning, that would allow people in the room to focus their questions on various aspects, but he didn’t do it. So we got a half hour into his presentation, and someone asked it again, and basically he ignored the question. Finally after a lot of people around the room had asked questions that would have gotten at it . . . I asked the question a little bit more specifically. He said he didn’t bring the data and couldn’t present it. That was exasperating and for the first time gave me the feeling that either the data didn’t exist at all or there were major restrictions on what he was willing to release in terms of information.” After the two seminars and yet another press briefing, Pons met with the director of Los Alamos, Siegfried Hecker. They discussed the possibility of a collaboration, and Pons seemed agreeable. If nothing else, the Utah chemist said he needed an inexpensive source of heavy water. He asked whether Los Alamos could get it for him cheap. Hecker assured him that would be no problem.

47

PALO

ALTO

In the past ten years, Joel Shurkin, a public relations officer at Stanford University, had organized three, maybe four press conferences. ““You have to win the Nobel Prize to get one out of me,” he said. On April 18 he had three in one day. Two could be explained because not all the press could get to Stanford in time for the first one. And Shurkin liked to separate the written press from the television people, which explained the third. It was all for cold fusion. Robert Huggins, said the official Stanford announcement, a distinguished professor of materials sciences, had run “‘the comparison that the physics community insisted upon.” The results were consistent with those of Pons and Fleischmann. Huggins and his researchers had run a heavy water cell versus a light water cell, and the heavy water cell, quite

226 8 GARY

TAUBES

simply, had run hotter. The effect was “real and substantial,” and they had confirmed it repeatedly. “Something is happening,’ Huggins said. “We don’t know what it is.” He added that they had eliminated any possibility that it was “some weird chemical reaction,”’ which seemed to leave only some

less weird

chemical reaction or some nuclear reaction. Huggins would not go quite that far. “We have a lot of ideas, but we don’t believe any of them,” he

said. ‘There are those who have no ideas and believe all of them.” Like those of MIT or Harvard or Caltech, an official Stanford Univer-

sity announcement is not something to be taken lightly. Huggins, in fact, quickly took to referring to himself in cold fusion meetings as “HugginsStanford,” as though it were his patronymic. (As Jim Brophy would say after the Utah claims had been rocked by serious criticism from MIT, “Pll see your MIT and raise you a Stanford.” So the point was not lost on Brophy either.) But Stanford was no more monolithic than MIT, and Bob Huggins was another one of the peculiar enigmas of cold fusion. He was the son of Maurice Huggins, who had been a world-renowned chemist. In his youth, Maurice Huggins had collaborated with Linus Pauling, and his work was considered by many chemists to be worthy of the Nobel Prize, although he never won it. Bob Huggins, who had just turned sixty, had founded the materials research department at Stanford a quarter of a century earlier. He had then spent two years in Washington running the materials research division of the Defense Advanced Research Projects Agency, DARPA. “‘At that time,” Huggins said, “‘the student protests were beginning to erupt, this was 1968. And it was beginning to have a very large impact on the Washington scene. There was a very serious possibility that the Department of Defense, who was funding us, would turn off all this research. So I took that job essentially to defend the program, which I was able to do.” In the years after Huggins returned to Stanford, his laboratory was flush with funds from DARPA,

the Office of Naval Research, and the

air force. In the 1970s Huggins built his materials science lab into perhaps the most prestigious in the nation. He became one of the key figures in a discipline known as solid state ionics, in which solids that readily diffuse ions are used to do a variety of chemical measurements and synthesis. His early papers are still considered the most rigorous and comprehensible in the field. Huggins had quite a few good ideas, and he lured excellent postdocs, who later said his ideas always needed filtering. In the five years preceding cold fusion, Huggins had not had it easy, however. He had suffered through a divorce and the death of his father,

BAD

SCIENCE

@ 227

and his professional edifice had begun to decay as well.? The solid state ionics work had been taken as far as it could go. Materials researchers said that much of the recent work coming out of Huggins’s lab was variations on the same old theme. One of his former students described this as the “well-developed Huggins set of tools. They’re pretty predictable.” Since 1982 or so, Huggins had lost virtually all his graduate students and postdocs.* “Graduate students began to show reluctance to work for him,”’ explained Ian Raistrick, a Los Alamos researcher who had been

with Huggins from 1973 to 1984, ‘“‘because quite a few had left without completing their thesis. And slowly the word gets around the department that if you want to get a Ph.D. in the least painful manner, you don’t go work for him.” In the over four years since Steve Crouch-Baker had come from Oxford for postdoctoral work and become Huggins’s chief researcher, funding had fallen off to virtually nothing as well. In 1989, for instance, Huggins’s funding, from the Department of Energy, which was his primary source, was only $145,000 a year. This was barely enough to sustain a small staff, let alone do research. When Huggins read the news about cold fusion on March 24, the first thing he did was call Turgut Giir, his senior research associate. Giir had obtained his Ph.D. from Huggins back in 1976, then had returned to his native Turkey to work in industry. In 1987 he came back to the lab. Giir had intended to spend the long Easter weekend skiing and was already comfortably ensconced in Lake Tahoe. ““There’s this very important news from Utah,” Huggins reported. ““They have shown that they have achieved cold fusion in a small electrochemical

cell, and that’s

something we could do in various different ways with very simple materials, with the background and experience that we have. And it has very important implications if this were true.’”’ Giir was evidently less enthusiastic than Huggins; he didn’t return to campus until Monday. After the weekend,

Giir, Crouch-Baker,

and Martha Schreiber, an

Austrian postdoc, began pursuing cold fusion. It took them a week to gather the materials, the most difficult to obtain, of course, being the

palladium. They scrounged through their storeroom and emerged with an old palladium crucible. They arc-melted it to purge it of hydrogen from any previous uses, then cast it into disk-shaped electrodes. (Huggins later attributed his cold fusion result to the peculiar origin of the palladium or his arc-melted casting process. It seemed at times to be the only palladium in the world capable of igniting nuclear fusion at room temperature.) The Stanford group initiated experiments on April Fools’ Day. Their equipment, as Huggins told reporters seventeen days later, was “‘remark-

228 8 GARY

TAUBES

ably primitive,’ but after all they were engineers, not physicists or chemists. At first, the Stanford researchers borrowed Geiger counters to be on the safe side, but they soon ignored the nuclear radiation issue.

Huggins was not interested in looking for nuclear products. His theory appeared to be that if one didn’t look for them, they didn’t exist. “You'll notice that I didn’t say anything about neutrons,” he said. ‘“That’s wonderful, isn’t it? You don’t want neutrons. To the physics community, the conventional fusion community, if you don’t have neutrons, you don’t have fusion. On the other hand, that’s the worst part of fusion. So to have some kind of thing that is giving you energy without neutrons, you’re much happier.” Huggins seemed interested only in proving what he could prove. So Giir, Crouch-Baker, and Schreiber built two reasonably identical cells.

They filled one with light water and one with heavy water, and put them side by side in a bath of water in a picnic cooler. On April 4 they began running the twin cells. As they took temperature measurements, Huggins sent out for Chinese food or drove down in his Mercedes to pick it up. This, Huggins said, was his only role, along with directing the research. “I probably spent about a thousand bucks myself feeding all these guys,”’ he said. On Apmil 12, according to Huggins, they first realized that the heavy water cell ran hotter than the light water cell. The difference was between one and one and a half degrees. But it was hotter nonetheless. This coincidentally was the same day that Harold Furth, head of Princeton’s hot fusion effort, challenged Stan Pons to run just such an experiment. Thus, as Huggins told reporters, “the comparison the physics community insisted upon.” (This, of course, was more thanalittle disingenuous,

unless one believes that running controls is something only physicists consider worthy of effort.) Huggins promptly invited representatives of EPRI, the Electric Power Research Institute, to see his one-day-old results “because it was pretty obvious that if indeed this turns out to be an energy source, it should be of interest to EPRI.”’ This might have made Huggins the first individual outside Utah to try to use cold fusion to increase his funding level, although others would quickly follow. Now, less than a week after Huggins had let EPRI in on his secret, he did the same with the rest of the world. His experiment, he said, “negates any possible chemical effects.’’ The press release explained that “because heavy water, containing deuterium, a heavy form of hydrogen, and regular water, containing simple hydrogen, are almost chemically identical, they would produce identical results if the effect were caused by a

BAD

SCIENCE

§ 229

chemical reaction.” And then, Huggins said, “the results [in the heavy water] seem to be larger than any chemical reaction we know about can

justify.” The Los Angeles Times reported that the Stanford result was the most difficult for cold fusion doubters to explain away. ““When [Nate] Lewis was asked about the Stanford results in the Caltech seminar, he shrugged and said, ‘He [Huggins] is a good guy.’ ” Lewis was being political. He was one of the first to realize that something was very wrong with the Huggins results. Immediately after the Stanford announcement, Lewis called Huggins to get details of what he had done. Whatever it was, Lewis wanted his group to try it, but Huggins wouldn’t give him any of the details. Said Lewis, “‘He didn’t know any of the details.” What neither Huggins nor apparently any of his researchers knew was that there are indeed chemical differences between heavy and light water, especially once lithium is added, as it was in the Pons-Fleischmann electrolyte. This had been in the scientific literature since 1958. It seems the electrical conductivity of heavy water with lithium is considerably less than that of light water with lithium. And this difference is more than enough to account for the heavy water cell running hotter, so it is the natural explanation for Huggins’s results. In fact, Huggins and his researchers had not been the only ones to run the comparison that the physicists had insisted upon. For instance, this was the very first experiment that Chuck Martin at Texas A&M had run. Martin ran a D,O cell and an HO

cell, and the D,O cell ran some two

degrees warmer. As Dell Lawson, who was Martin’s graduate student, told it, “We went, “Wow, this is real great.’ And we were all very excited. But then we said, ‘We need to measure conductivities of these two solutions.’ And as it turned out, the conductivity of LiOD in D,O

is about a factor of 1.5 to 2.0 smaller. That means that the change in energy across the D,O cell is higher, so that means that more power is dissipated, therefore the temperature is higher. When we heard Huggins’s initial results, we thought, Gee, they’re making the same mistake we did.’’*> None of this information, of course, was in the Stanford press release,

nor did Huggins mention it to the reporters. Huggins may not have realized the error until July. In Utah, the administration had known about the Huggins confirmation since April 12. ‘‘For a while there,” said Hugo Rossi, “Huggins was our savior. He brought us over the hump.” Stan Pons certainly seemed to see Huggins as a savior. He immediately told reporters that the most important confirmations were from Huggins at Stanford and Martin at

230 8 GARY

TAUBES

Texas A&M (although Martin already knew that his results had been compromised). In fact, immediately after the Huggins-Stanford confirmation hit the press, Pons told The Salt Lake Tribune that the only electrodes that would induce cold fusion appeared to be those that were cast, as opposed to extruded or machined. Pons had never before claimed that his electrodes were cast, or at least that claim hadn’t shown up in print. Fleischmann and Hawkins had both said that they used the electrodes as they came out of the box. Huggins, however, had cast his electrodes; now Pons began

insisting that cast electrodes were the only ones that worked; he now knew this to be a fact and could reproduce the excess heat effect 90 percent of the time.

“I think it’s clear now,”

processing of the palladium may be This piece of misinformation then It sent all the aspiring cold fusion electrodes, or casting their own in done until then had been useless.

Ce

With the news

ee

Pons

said, “‘that the

afactor.” meandered through the rumor mill. researchers out searching for cast the belief that everything they had

Ce

out of Stanford, the situation, as one Department

of

Energy official put it, ‘“had come to a head.’’ The department had had its laboratory administrators send emissaries to Washington immediately. On the morning of Apmil 19, the various administrators met with Louis Ianniello, the deputy director of the Office of Basic Energy Sciences at DOE, who told them that James Watkins, the secretary of energy, had made the pursuit of cold fusion the department’s highest priority. He wanted a written report every week summarizing the research. The government laboratories had free reign to pursue their cold fusion research, Ianniello said, to use whatever resources

they needed,

and

DOE would cover the expenses. He emphasized that the DOE brass wanted no secrecy between the labs themselves or between the labs and DOE headquarters. The message was clear: the DOE administrators did not want to be surprised by anything they read in the press. If one of their labs had the slightest sign of cold fusion, they wanted to know immediately.

BAOMSCIE NCE: 1.231 49

SALT

LAKE

CITY.

PROVO

é

Once Peter Hagelstein of MIT had gone public, the cold fusion interests at Utah had begun mobilizing to protect their property. Simultaneously, they heard about the confirmation from Huggins-Stanford. While Huggins may have appeared to be the savior of cold fusion, his results also made him, and Stanford, a prime competitor for patents and rights. Paul Van Dam, the Utah attorney general, announced that he had hired the firm of Giauque, Williams, Wilcox & Bendinger to protect cold fusion for the state. Since all four were trial lawyers rather than patent lawyers, Richard Giauque explained that he was the “point guard” on the cold fusion legal team. His first task was to hire the best patent lawyers and pass them the ball.*¢ At this time Norm Brown and Peter Dehlinger had been running the cold fusion patent show for the university. Brown was doing with cold fusion what he’d done with the other 121 inventions he’d overseen at the Office of Technology Transfer. Along with his boss, Jim Brophy, and Chase Peterson, Brown had formulated some general principles, the foremost being that Utah would not exclusively license cold fusion to any one major player but, rather, would make the technology available to all interested parties.°”? Brown was prepared to charge firms $20,000 to option the cold fusion technology. In the month since the announcement, he’d had some 200 inquiries from private enterprises. Of these, 65

signed confidential disclosure agreements allowing them to view the patents. (Nate Lewis of Caltech also tried to sign a nondisclosure agreement to see the patents, but, not surprisingly, he was not allowed to do so.) Twenty of these were willing to pay $20,000 to option the technology, but none ever would. When Richard Giauque came on board, he insisted that cold fusion was too big a project to treat like any other university invention, and he considered Brown and Dehlinger minor leaguers who didn’t understand the game of hardball as it was played in the big leagues. (Dehlinger later described Giauque as a “‘slick Salt Lake City trial lawyer” and agreed that by Giauque’s standards he was small-time.) Both Giauque and Ian Cumming, who was now running the cold fusion campaign with Peterson, considered Brown’s idea of raising fusion money through options the equivalent of “wearing sandwich boards and selling tickets.” If Brown could be so misguided on this point, then they had to wonder about the quality of the patent applications. So Giauque set about getting Brown

232 BU GARY

TAUBES

off the cold fusion project, running Dehlinger out of town, as Dehlinger put it, and searching for the best patent attorneys money could buy. By this time the BYU administration had also set out to protect themselves, with three goals in mind. The first was to establish the provenance of the electrochemical fusion, which they believed would be Steve Jones. The second was to clear the air of these piracy accusations and clear Jones’s good name. The third, as Lamond Tullis said, was to “protect our own research territory so that we wouldn’t be proscribed at some point in the future of being able to pursue our own research.” At one point the BYU administration even considered suing the University of Utah for defamation of character regarding Jones and the piracy accusations, but they decided it was the ugly alternative. Instead, they would apply for their own cold fusion patents, assuming that this would prompt what are called interference proceedings in the patent office. These proceedings would then force Utah to open their books and prove that Jones was at least innocent of piracy, if not the original discoverer of electrochemical fusion. Said Tullis, “We really engaged in

a vigorous patent activity.”’ He added that they did so out of righteous indignation and “‘not from the vantage of capturing a whole pile of money, although somebody may make a pile of money out of this.” The board of trustees of BYU gave the administration carte blanche to spend what it would take to protect themselves. Tullis spent the early part of April interviewing the premier patent firms in Washington and New York, many of which had to explain to Richard Giauque when he came calling that they couldn’t take his business because they had a conflict of interest on their hands. Eventually Giauque hired the Houston firm Arnold, White & Durkee, considered

to be one of the better patent firms in the country. Between them and the local lawyers, Giauque, Williams, Wilcox & Bendinger, plus C. Gary

Triggs, who was working for Stan Pons but billing the university, the attorneys figured on spending the entire $500,000 allotted to them before the end of the year. Giauque and his colleagues would charge $34,000 a month through November, which wasn’t bad considering the firm didn’t specialize in patents. The Houston firm would charge $190,000 by the end of October. Both firms, according to the Utah attorney general’s office, were billing at reduced rates, and both of them based

their estimates on the assumption that they would not be, as The Salt Lake Tribune reported it, “dealing with ‘major’ litigation by year’s end.”’ Joe Tesch, the Utah assistant attorney general, told The Deseret News that the Houston firm was necessary, if nothing else, for intimidation. ‘This is a psychological thing,”’ Tesch said. ‘““To be quite honest, Utah

BAD

SCIENCE

&@ 233

is not taken very seriously by a lot of people. We needed to elicit the biggest bulldog on the block to let people know how serious we are, to create a national and international impression, and to let people know not to mess with us.” While the various lawyers set out to protect Utah’s legal right to cold fusion, the Washington, D.C., lobbying firm of Cassidy & Associates would concentrate on protecting and furthering the U’s cold fusion interests in the nation’s capital. Cassidy & Associates had made a lucrative business obtaining funds directly from Congress—a process known as earmarking, or, more colloquially, “‘pork,” short for pork barrel—for a variety of university and medical research facilities around the country. For their services, the firm charged $30,000 a month, with a two-year

minimum.** Chase Peterson didn’t consider $30,000 per month exorbitant. “If you ask what is the Washington lobbying cost for the University of Illinois,” he said later, “‘I’ll bet it’s a heck of a lot more than $30,000

a month. They probably have four people full-time in Washington. The University of California has got ten, and we don’t have anybody.” The U also recruited Ira Magaziner, whom the Tribune described as a $1-million-a-year business consultant, who had consulted for Ian Cumming in the past. On paper, Magaziner appeared to be a cross between Ralph Nader and an Ayn Rand-style superman. He had been acollege activist at Brown University and still received much of the credit for revamping that school’s curriculum.®? He was a Rhodes scholar; a cofounder of Telesis, an international consulting firm; coauthor of Minding America’s Business; and author of The Silent War: Inside the Global Business

Battles Shaping America’s Future. Magaziner lectured, consulted, and wrote about achieving social and economic progress while preserving the environment and beating the Japanese in what he thought of as the Technology War. On the night of April 17, Magaziner flew into Salt Lake City on Cumming’s private jet. He met with Peterson and company for breakfast on the eighteenth and elucidated what Peterson later called ‘‘this highfalutin Ira Magaziner thesis,’ which happened to resonate nicely with his own: “This country is going down the tubes in its productivity. Practically every project of which the basic science has been done in America is now being produced elsewhere, from color television to microwaves to superconductivity, to high-definition television, to supercomputers. If there is any merit to cold fusion, can we afford to let it go through the very slow and methodical process of development and substantiation; and after months and years, then engineering studies to see if something useful is to be gotten out of it, then after that, to see if there is any

234 BE GARY

TAUBES

commercial way to use the engineering ideas? That is the serial process of American science, and it’s failed. Magaziner’s thesis, and I buy it, is we

need to truncate that and make it parallel. We must be willing to invest in the engineering development of ideas even before they’re proved. And you say, ‘Well, that wastes money.’ And the answer is yes, it does. But you’re also going to have some winners. The Japanese are doing it, and we thought we should too.”’ The Japanese quickly became to cold fusion what the Soviets had been to the defense community, which is to say the Evil Empire. If the United States didn’t do it now, the Japanese would do it first and better. This argument had some logic to it, although the idea of developing technology based on an unproven and perhaps incorrect scientific principle is a highly debatable one. Still, whenever Magaziner appeared in the cold fusion issue, he was followed by a cloud of horror stories describing the concerted Japanese effort to appropriate cold fusion for the good of Japan.

On April 18, the day Magaziner appeared publicly on the Utah bandwagon, The Wall Street Journal ran the first of what would be many Japanese-as-Evil-Technological-Empire stories in cold fusion: Toxyo—Japan has joined the race to verify the University of Utah report that nuclear fusion can be accomplished in a laboratory flask. . . . The scene [in Japan] is reminiscent of Japan’s quick action two years ago after University of Houston researcher Paul Chu unveiled a new hightemperature superconductor. The Japanese quickness to pick up on the new technology alarmed some Westerners. As companies filed thousands of superconductivity patents, engineers boasted that Japan would probably walk away with the commercial applications. It’s much too early to speak of a Japanese challenge in cold fusion. But it’s clear that Japan has caught, as one Japanese researcher calls it, ‘fusion fever.”’

A year later, Magaziner would tell The Providence Journal that his involvement in cold fusion had been suspended because he still didn’t “know if there is anything real or not here, and since it is clear that the economic uses are a long way off.’’ He then added, nonetheless, that “‘in

Japan, it’s viewed as a great breakthrough.”’ Those Japanese were beyond comprehension.

BAD

SCIENCE

§ 235

SO ROME, FRASCATI, HARWELL, SAO PAULO, DRESDEN, SEOUL, “BEIJING,

AND

GAINESVILLE,

FLORIDA

After Huggins-Stanford, it was the Italians. On April 19, representatives from

the Italian Commission

for Nuclear

and Alternative Energies,

ENEA, appeared on Italian television standing in line at the Rome patent office, waiting to officially claim their share of the cold fusion future. This transaction was timed to occur shortly before a press conference held by Francesco Scaramuzzi of the Frascati Energy Research Center, which was run by ENEA. Scaramuzzi had placed one hundred grams of titanium shavings in a canister, pumped in deuterium gas, then submerged the canister in a thermos full of liquid nitrogen. This approach had not been published by either Utah or BYU and was thus patentable as far as the ENEA was concerned. But it was not new—it was an inefficient variant of the vacuum-loading technique that Paul Palmer had used at BYU in 1986. ENEA claimed it was ‘‘simpler and more direct”’ than bothering with electrolysis and electrolytes, which it was. It became known as “‘dry fusion.” Scaramuzzi ran the experiment twice: once between April 7 and April 10, which must have been the source of the rumors that had circulated at the Erice conference, and the second time on the fifteenth

and sixteenth. Both times his single BF, detector appeared to detect bursts of neutrons. Between the two runs, Scaramuzzi and his colleagues accidentally destroyed their cell, thus the five-day hiatus. During the first run, the neutrons appeared in bursts of 20, then 40, then 20, then 40. Even Scaramuzzi found this disturbing enough to suggest that the neutrons had saturated his detector. The second time around the canister seemed to be emitting as many as 1000 neutrons per hour—500 times the background—and the signal rose slowly over the first seven hours, then decayed away. Scaramuzzi must have realized this was not a definitive experiment. But in what was becoming a familiar scenario, the Frascati

physicists later said that they had gone public because other scientists in the lab had seen their results. The Italian news media reported that ENEA had confirmed and ‘“‘several years would be needed to turn the laboratory success into the energetic revolution everybody waits for.” Obviously the local journalists were not familiar with the idiosyncrasies of BF, neutron detectors. Scaramuzzi seems to have been less than

familiar with them himself. When he took to showing his data later,

236 B GARY

TAUBES

those physicists and engineers who had worked with the neutron detectors immediately recognized the signs of electronics problems. Whatever Scaramuzzi had seen, it almost assuredly wasn’t neutrons from his cold

fusion canister.°° Harwell tried to duplicate the Frascati results using helinm 3 detectors, which are considerably better than BF, detectors. In any case, they set up the experiment, complete with the liquid nitrogen, and went out to lunch. They came back from lunch, said Martin Sené, ‘‘and there was on

the chart recorder this great big peak. Neutrons everywhere.’’ However, they had two detectors, and only one had observed the neutrons. This was curious. Sené discovered that he could apparently generate as many neutrons as he liked by putting his foot against a nearby electric socket, which probably caused a spark, which set off the detectors. He speculated that as the canister warmed up from the liquid nitrogen temperature, water vapor was produced, which led to shorting in the electronics. “We rerouted all the electrical cables,’ he said, “‘and we never saw anything

like it again.” On April 19 researchers at the Institute of Energy and Nuclear Research at the University of Sao Paulo in Brazil reported that had they passed an electric current through a one-millimeter-thick palladium sheet immersed in a bath of heavy water. Lo and behold, the neutron radiation doubled when the current was on. Spero Penha Morato, chief of the special processes department at the institute, said, “I would tear up my Ph.D. if it is not a nuclear reaction.”’ The press reports did not say what kind of neutron detectors Morato used, nor what he later did with

his doctoral diploma. The wire services reported that researchers from Dresden in East Germany had also confirmed and that Dr. Yoon Kyong-Sok of the Korea Advanced Institute of Science and Technology and Dr. Lee KyuHo of the Korea Research Institute of Chemical Technology had observed cold fusion in separate flask experiments. For this, according to the AP-Dow Jones News Service, the Ministry of Science and Technology “will propose the government invest $1.5 million in cold fusion research)” In China the Xinhua News Agency reported that “leading physicists and chemists from a dozen research laboratories have attempted to recreate the room-temperature nuclear fusion process” and been unable to do so.° And, finally, the University of Florida announced that two retired nu-

clear engineers, Glen Schoessow and John Wethington, had produced

BAD

SCIENCE

§ 237

tritium by the test tubeful using a homegrown variant of Pons and Fleischmann’s technology. Schoessow and Wethington charged acell for forty-eight hours, according to the university press release, after which they detected one trillion atoms of tritium, and that number increased another twentyfold

after a hundred hours. They ran control experiments “using heavy water without electrolysis,” which implies that they compared the experiment with a glass of heavy water. The press release also claimed that Schoessow and Wethington had specially treated their palladium before beginning the experiment, although Schoessow wouldn’t say how, and that they ran above room temperature but below the boiling point: ‘““Schoessow said, that if the [University of Florida] cell indeed produced fusion, it would be ‘warm fusion’ rather than the ‘cold fusion’ others have been trying to achieve.” If one is of a cynical turn of mind, this last statement might sound suspicious. It seems like unnecessary quibbling to claim the discovery of yet another heretofore unknown physical process on the basis of a few tens of degrees in water temperature, when the bench mark is the several million degrees required for hot fusion. Apparently Schoessow and Wethington, tritium or no tritium, intended to stake a patent claim that would cut them in on a piece of the warm fusion action. Later they graciously refused invitations to present their data at conferences, or requests from anyone to see their experiments, except, apparently, Martin Fleischmann, claiming that their pa-

tent lawyers advised against it. Nate Lewis, as was his habit, called Schoessow and Wethington after

reading about their discovery in the Los Angeles Times. They told him what they had done, and he remarked that there was a common artifact known as chemiluminescence, which would fake a tritium signal in the

tritium counter they were using. The two Florida nuclear engineers professed to be ignorant of the effect, which seemed unfortunate, albeit

typical of cold fusion, considering they’d already gone public. Luis Muga, a professor of chemistry who worked upstairs from the two, later remarked, ‘““They’re very kindly gentlemen, real nice people.” He added that Schoessow was an octogenarian; Wethington was a few years away, and the tritium work was a “‘very large blunder.” Muga believed his friends were not as critical of their experiment as they might have been. When he heard about the press release, Muga asked a health physicist to check out the heavy water supplies in the nuclear engineering lab, and they did have a relatively high concentration of tritium. As

238 # GARY

TAUBES

Muga later reported, the health physicist also observed that their tritium counter was not properly calibrated. Four months later Schoessow and Wethington told Utah officials that they could produce gamma rays, tritium, and heat on demand, and enough energy to illuminate a light bulb. One year later Chase Peterson said that two Florida scientists had offered to sell their special fusion formula to the University of Utah for $1 million. This sounded like the retired warm fusion duo, and it was. As Hugo Rossi recalled it, Schoes-

sow told the Utah officials “that he really understood something that nobody else understood, and was willing to transfer those patents for a million bucks. He argued that since he was in his eighties, a million dollars meant the same to him as ten million or one hundred million.” By this time, the cold fusion proponents at Utah were desperate. Still, Peterson passed up this spectacular offer.

51

NATURE

John Maddox was a sixty-four-year-old, chain-smoking, Oxfordeducated physicist who left academia to work for the Manchester Guardian and left the Guardian in 1966 to edit Nature, a job he had been doing off and on ever since. When asked what he considered Nature’s function to be, Maddox replied, ““We are a trade magazine. We stand in relation to the scientific community as does the American journal Casket & Sunnyside in relation to morticians in the United States.”” The response reflects both Maddox’s dry wit and his belief that his 130-year-old journal should be a wide-ranging forum for scientists. He published not only scientific papers but comments on the manners, morals, and conduct of the reader-

ship. These had put Nature in the news almost as frequently as some of the seminal papers it published. In July 1988, for instance, Nature published an article by Jacques Beneveniste, a French chemist and pharmacologist. Beneveniste claimed that a particular human antibody could provoke an allergic reaction in white blood cells, even when diluted in water to the point—one part in 10'2°—that not a single antibody could conceivably remain. In effect, Beneveniste was claiming that water could be imprinted with the re-

BAD

SCIENCE

# 239

membrance of things past, which smacks of homeopathy. As it turned out, Beneveniste was funded by a French maker of homeopathic medicine. Beneveniste had badgered Maddox for two years to publish his paper, protesting that he was suppressing “‘the greatest discovery of the century.”” Maddox finally agreed, on the condition that the experiments be repeated in the presence of an on-site investigative team. Beneveniste accepted the challenge, and Maddox, Walter Stewart, an expert in fraud from the National Institutes of Health, and James Randi, also known as

the Amazing Randi, a professional magician and amateur fraud buster, went off to France. The three sat in on a week of experiments, during which Beneveniste was unable to reproduce his effect. They then took photographs of 1,500 lab-book pages home with them. Maddox published the conclusion of his investigation, reporting that Beneveniste’s claim was “‘as unnecessary as it is fanciful.” Beneveniste responded that he had been the victim of a “Salem witch-hunt.” He warned his peers to “never let these people get in your lab,” which made him sound like Henry Higgins. He insisted that he was being mocked for his great ideas and likened himself to Galileo. All this had little impact on Maddox, who believed that ‘“‘somebody who has made a very earth-shaking discovery must expect to spend many years trying to make this come true, convincing people, producing the evidence.” It was all part of the screening process. Maddox,

in turn, was criticized for debasing his journal, and even

science, ““‘by washing dirty linen in public.” He responded with a letter to The New York Times, arguing that the well-being of the scientific community depends on the recognition “‘that second-rate science exists, can be exposed and should be more openly categorized as such.” When Maddox read about cold fusion in the Financial Times six months later, his first reaction was “Gee whiz.” He quickly became dismayed when he read that the paper had been sent to Nature. “We hope that if people are going to ask us to print their papers, they ought to let us have them before they tell the world.” Ordinarily, Maddox explained, he would have been indignant and refused to publish on principle: ““We have a rather strict rule that people shouldn’t talk about papers before they’ve been published. But this is obviously a matter of enormous public interest.’’ Still, Maddox could detect, yet again, an aroma of canard hanging over the story. He sent the Utah and BYU papers out for peer review and while waiting for the critiques, wrote a wry, vitriolic editorial on the importance of peer

240 © GARY

TAUBES

review before public announcement of new results. He entitled the editorial CoLD (CON)FUSION. It argued that “reports that an account of cold nuclear fusion is soon to appear in this journal are premature”: It is the rare piece of research indeed that both flies in the face of accepted wisdom and is so compellingly correct that its significance is instantly recognized. Authors with unique or bizarre approaches to problems may feel that they have no true peers who can evaluate their work, and feel more conventional reviewers’ mouths are already shaping the word “‘no” before they give the paper much attention. But it must also be said that most unbelievable claims turn out to be just that, and reviewers can be forgiven for perceiving this quickly. . . . But patents that turn out to be worthless and investments that disappear may demonstrate the value of cautious peer-review to those who now think of it as a fusty institution much loved by pedants.

The editorial prompted The Deseret News to suggest that the state legislature might not fork up $5 million for cold fusion if Nature did not publish the Utah paper. Bud Scruggs, the governor’s chief of staff, then responded, “‘We are not going to let some English magazine decide how state money is handled.” On April 20, Maddox announced that Nature would not be publishing the Utah paper because all three of the referees had serious criticisms of the work, but they would publish the BYU paper in the next issue. He then explained that Pons and Fleischmann had simply not done the ‘rudimentary control experiments of running their electrolytic cells with ordinary rather than heavy water.’’ He added that scientists had every right to be indignant at this neglect: How is this astonishing oversight to be explained to students repeatedly being drilled in the need that control experiments should be as conspicuous in the design of an investigation as those believed to display the phenomenon under study? And how should the neglect be explained to the world at large? There is no convincing explanation, only extenuating circumstances. Self-imposed secrecy has evidently hampered the investigators, understandably buoyed up by their belief that they had discovered a remarkable new phenomenon and fearful that too much talk about it would give other bigger battalions a chance to steal a march on them. Yet it is unthinkable that, if the authors had felt able from the outset to stand in

front of routine laboratory colloquia and give a full account of their work, the question “Have you tried it with ordinary water?”’ would not have

BAD

SCIENCE

@ 241

been raised. This glaring lapse from accepted practice is another casualty of people’s need to be first with reports of discovery and with the patents that follow.

All in all, Maddox wrote, “The Utah phenomenon is literally unsupported by the evidence, could be an artefact, and given its improbability, is most likely to be one.’ When Maddox published Steve Jones’s paper—thus fulfilling the worst possible scenario foreseen by the Utah contingent back in early March—he wrote that this should not imply that “the experiments described elsewhere by Fleischmann and Pons are inherently less believable than those of Jones,” and that the appearance of the Jones paper “should not be taken to imply that all those who have seen it are persuaded to its chief conclusion.’ To support this point, he simultaneously published one of his referee’s critiques, which suggested that the BU signal could likely be explained by natural fluctuations in the background rate of cosmic rays.* The University of Utah responded to the rejection of the PonsFleischmann paper with its increasingly familiar spin. Pam Fogle told The Deseret News that Nature “specifically asked for more data and to flesh out the article by some 1,000—2,000 words. Some of the data they were asking for was coming out of experiments that are currently operating. [Pons and Fleischmann] didn’t have the data to supply to the reviewers for the second go-round.”’ Since Nature had asked, among other things, for data concerning controls, that still begged the question of whether the controls had been done. Perhaps these were the experiments Pons was operating on April 20, nearly a month after the announcement (although Mark Anderson, who did the bulk of the cold fusion work in Pons’s lab

from March 25 through the middle of August, recalled that they only began running light water controls in the middle of June). After the Maddox editorial, Pons told reporters that he “‘might’’ not agree that plain water is a control experiment. “I'll argue,” he said, “‘that the control is taking a palladium rod that absolutely does not work, and place that side by side to an experiment that is, in the same solution, and that is a control experiment. It’s never done anything but sit there and look at you.’’ How this constituted a control is a little unclear. Apparently Pons considered it a test of his calorimetry. If an electrode that did not produce excess heat never produced excess heat, then his calorimetry was accurate. Maybe so. It still begged the question of whether fusion was responsible for the excess heat in a ‘working rod,” which a light water test, run correctly, might have answered.®

242 © GARY

TAUBES

Meanwhile, Pons was informed that researchers from Drexel Univer-

sity in Philadelphia had run light water against heavy water and both had given off excess heat. Thus, they concluded they were dealing with a chemical reaction. To this news, Pons replied in The Deseret News, “They don’t have to believe me. I will just go back to. the lab, do my experiments, and build my power plant.”

52

THE

PRESS

Partial responsibility for Stan Pons’s blind faith in the existence of cold fusion has to go to the rarefied environment in which he was living. It was hard not to believe in cold fusion in Salt Lake City. The two local newspapers, The Salt Lake Tribune and The Deseret News, not to mention The Wall Street Journal, were all, to varying degrees, pro—cold fusion. By approaching the story as though it were a political election or a sporting event, the editors of these papers seemed to believe they could help the home team triumph. Whether this was compatible with good journalism or the needs of their readership is subject to debate. The Deseret News began running its articles under a logo proclaiming UTAH’S FUSION FUTURE. This was soon accompanied by a daily “Fusion Scorecard,”’ which listed the status of national and international cold

fusion experiments under the misleading categories ‘‘confirmed” and “collaboration.” On April 21 it read: Fusion Scorecard Confirmed U. experiment: Lajos Kossuth University, Debrecen, Hungary Texas A&M University Moscow University

University of Washington Italian National Agency for Alternative Energy Stanford University Czechoslovakia team of physicists University of Florida, Gainesville Collaboration: Los Alamos National Lab India scientists

BAD

SCIENCE

§ 243

The Deseret News acknowledged that cold fusion had its skeptics, but they were, of course, “mainly physicists who have chased nuclear fusion for decades.” The slant of the News may have been accentuated by the fact that for quite a while its science reporter, JoAnn Jacobsen-Wells, was the only local reporter Pons would talk to,® and she seemed to treat cold fusion as if it were the proverbial manna from the heavens. The editors of The Deseret News openly acknowledged that their reporting might be one-sided, but they believed they were serving the good of the people. Editor Lavarr G. Webb wrote in an editorial: While the Deseret News hasn’t exactly pulled out the headline we're saving for the Second Coming, | have to admit we did get excited about this story. . .. We don’t want to indulge in gee-whiz, cheerleading journalism. But, my goodness, when you have a number of the state’s leading scientists saying this may be the discovery of the century, the answer to the world’s energy needs, you can bet we’re going to report that with big headlines.

Webb explained their editorial position by arguing that they were, after all, a Utah paper and not the “good, gray, East Coast, nationally oriented New York Times.’’ He did promise, however, that “if the whole thing fizzles and becomes an embarrassment for Utah, we’ll cover that just as aggressively.” The Salt Lake Tribune, while considerably more restrained, still agreed that the critics of cold fusion were ‘‘mostly plasma physicists who have based their life’s work on the notion that fusion can occur only in an environment similar to the center of the sun.” But even these physicists, the Tribune reported, “‘have begun to accept the idea that a little fusion is occurring in the experiment.”’

53

CAMBRIDGE.

GAMMA

RAYS

CONTINUED

Four weeks after the Utah announcement, physicist Richard Petrasso of MIT still found that many of the scientists he spoke to about cold fusion would cite Pons and Fleischmann’s gamma ray spectrum as proof that cold fusion must be real. When Petrasso explained that the spectrum was assuredly spurious, they directed their skepticism at his motives rather

244 0 GARY

TAUBES

than at the quality of the Utah science. Exasperated, Petrasso decided to prove that the gamma ray evidence was bogus. “It was time to act, jhe said. The principal question was why Pons and Fleischmann had taken a spurious signal—obviously the product of an electronics glitch in their detector—and identified it in their Journal of Electroanalytical Chemistry paper as a gamma ray peak of exactly the energy that it should have had if it were real: 2.22 MeV. It might have been a coincidence that the spurious signal had actually appeared at the point on the spectrum corresponding to 2.22 MeV, but the odds against that were astronomical. The first place to look for an answer would be in the entire gamma ray spectrum itself, which Pons and Fleischmann had not shown in their JEAC paper, but which had been broadcast on television when reporters had interviewed Pons and Fleischmann in their lab on March 23. Stan Luckhardt, one of Ron Parker’s physicists, while watching CNN, had noticed the spectrum displayed on a computer screen in Pons’s lab, and he had given Petrasso a videotape. Petrasso froze the film when the camera panned up to the spectrum, then photographed the frame off the television screen. (Pons later told The Deseret News that this spectrum was not the real one but only a ‘““dummy”’ borrowed from another department on campus.) With the entire spectrum, Petrasso had Pons and Fleischmann’s spurious gamma ray peak, if he could identify it in the spectrum, as well as all the background peaks, which would make that identification possible. This background radiation comes from naturally occurring radioactive sources, which are continually and indiscriminately launching gamma rays into the environment.

Potassium 40, for instance, exists in trace

amounts in all glass—beakers, windows, sophisticated test tubes—and it tends to capture an electron and emit a gamma ray of exactly 1.46 MeV. (Although, to be precise, the potassium 40 actually decays first into argon 40, and it’s the argon that emits the gamma ray.) One physicist pointed out on the computer network that the “‘no-salt’’ salt substitute from the local grocery store comes with enough potassium 40 to calibrate a gamma ray detector “without even needing to fill out a radiation permit!” Then there’s thallium 208 hiding in concrete, which means it’s in the floors, walls, and ceilings of most buildings; it emits a gamma ray of exactly 2.615 MeV. The thallium comes from thorium 232, which has a half-life of 14 billion years. This, Petrasso noted, “‘is pretty incredible when you think about it. That’s almost the age of the universe.’’ Thorium 232 can come from the decay of uranium and itself decays in small

BAD

SCIENCE

8 245

explosive steps until it finally hits thallium 208, all of which will emit distinctly energetic gamma rays that show up in any spectrum. “‘I have never seen a spectrum in my life that didn’t have them,” Petrasso said. If Pons and Fleischmann’s gamma ray peak at 2.22 MeV had been real, it would have appeared in their spectrum nicely set off between the potassium 40 peak at 1.460 MeV on the left and the thallium 208 peak at 2.615 MeV on the right. 9

=

Cee

Y

=

if (1.460)

@

Dominantly

a

ggBie4

5

(1.764)

3

93Bin 4

{Baye

aula

a

(2.615)

0 A

oa

1.4

1.9

2.4

2.9

Energy (MeV)

Because Pons and Fleischmann had not bothered to show the spectrum, many physicists assumed that they had simply misidentified their peak. The expert on salt substitutes speculated that Pons and Fleischmann had picked up and misidentified a polonium 214 line at 2.204 MeV. Polonium 214 is also in concrete and comes from the same line of descent as thallium 208. Steve Koonin of Caltech speculated publicly that Pons and Fleischmann may have simply detected the dreaded radon line, but he admitted later that he wasn’t enough of an experimentalist to have picked up on the other problems with the gamma ray peak. When Petrasso examined the Pons-Fleischmann spectrum, he saw that

the peak, or the artifact, depending on one’s point of view, was at not

246 # GARY

TAUBES

2.22 but 2.5 MeV. He knew this because the spectrum clearly gave him the potassium 40 peak at 1.46 MeV, and he could measure upward from there. In fact, no peak could be found in the spectrum at 2.22 MeV,

which means there had been no gamma rays, and thus no neutrons. There was still the question, however, of whether the gamma ray spectrum that had been on television was the only one that Pons and Fleischmann had. If they had more than one, Petrasso’s evidence would be less than ironclad. To be sure of his data, Petrasso sent his student

Kevin Wenzel out to Salt Lake City to visit one of the local television stations. Wenzel spent an hour scanning film clips of the laboratory and returned with the news that only one spectrum existed. The logic of what Petrasso wanted to prove didn’t require that he have the Utah spectrum or even show it. But he wanted to show the scientific community that Pons and Fleischmann did have the entire spectrum, and that they could have shown it had they so desired. Step two was to set up at MIT a dummy Pons-Fleischmann experiment, or a pube (pronounced “‘poo-bee”’) experiment, as Petrasso called it. He and Wenzel put a plutonium -beryllum source, hence pube, in a water bath. The pube source would mimic a working fusion cell, if such

a thing existed, and emit neutrons. The neutrons would pass through the glass of the cell and interact in the water bath, releasing a gamma ray, and that would be observed by a detector identical to that used by Pons and Fleischmann. The detector observed a peak at 2.22 MeV, which meant that if the Pons-Fleischmann cell had been emitting neutrons, their

detector should have registered them as well. As far as Petrasso was concerned, this experiment was actually irrelevant, because they already knew exactly what such a spectrum had to look like. But he wanted to publish a letter in Nature with all the hard facts on the gamma ray spectrum, and he didn’t want anyone who read it to take anything on faith. “We wanted to do the experiment to show people what this peak has really got to look like.” Meanwhile Ron Parker tracked down Bob Hoffman, the health physicist in the radiation safety department at Utah who had done Pons’s gamma ray work. “It was very surprising,” Parker said, “because I thought that he would just clam up, because Pons was being very secretive, but he kept talking and talking. And by the end of the conversation I realized that he was really unsure about what they were publishing. And they had taken data from him, . . . never acknowledged him, and published the damn thing.” Hoffman did not say much about the data, but he did tell Parker that

BAD

SCIENCE

#8 247

he suspected that he had had instrumental problems. He wanted to set the equipment up again, but he was having trouble getting back into Pons’s lab. He added that he had given Pons the data tapes, and that he never saw the peak itself or discussed the data with Pons until it was published. He had spent a week and a half acquiring the data, but he would have liked another month. Between April 21 and April 26, Petrasso talked to Marvin Hawkins once and Bob Hoffman four times. According to Petrasso, both told him that they personally thought the peak was at 2.5 MeV and not at 2.22

MeV, and that they had told Pons that. (Pons had said to Hawkins, “‘T’ll

get back to you on that, Marvin,” but he never did.) Hawkins wouldn’t give Petrasso any of the raw data, because he said Pons wouldn’t want him to, he just confirmed facts that Petrasso already knew. Hoffman,

in contrast, apparently considered

Petrasso

a fellow afi-

cionado of radiation and read him the counts that had been recorded in each channel of the detector, and then told him which background peaks appeared in which channels. From that Petrasso re-created the spectrum himself, and it appeared that the only place the artifact could have been was at 2.5 MeV. With this information, Petrasso went to work on his letter to Nature,

which he was determined to finish before May 1, when the American Physical Society was holding a special cold fusion session at their spring meeting in Baltimore. His letter would prove dispassionately that the gamma ray peak could not be real, and certainly could not be at 2.22 MeV. Petrasso had trouble keeping to the neutral tone that scientists prefer. He later said that anyone reading the first draft of the paper would be shocked at how unscientific and emotional it sounded. His colleagues kept insisting that he rein himself in. ‘“They were really sitting there saying take that crap out,”’ Petrasso said. “‘Stick to the science. But it was really hard. It was like nothing I had ever done, read about, or experienced in my life.” Petrasso also fought with Ron Parker about citing the conversations with Hoffman and Hawkins as further confirmation of the peak being at 2.5 MeV. Parker said they had a shouting session over this: “‘I felt that too much of it was hearsay, and it was too easily denied. I wanted to stick to what we knew.” In addition, Hoffman and Hawkins had not provided any information that was critical to their argument, only corroboration. Parker won. (Of course, he said, “‘being director helps, you know.’’) So in the paper they only acknowledged ‘“‘M. Hawkins and R. Hoffman,

248 § GARY

TAUBES

personal communications” in those places that were secondary to their um on argument, for instance, where Petrasso claimed that the spectr

television was indeed the one and only spectrum from the lab. Parker had another motivation for not mentioning Hawkins and Hoffman any more than they had to: “T felt it was entirely possible,” he said, “given the way [Pons and Fleischmann] were acting, that they would just fire Hoffman. Furthermore, Hawkins was finishing his Ph.D. thesis, and we didn’t want to queer his thesis and destroy his chance of

getting a Ph.D. or anything.” Parker then insists that he even considered that if either Hawkins or Hoffman got fired, he’d hire them at MIT. “We just can’t do this to a person,” he’d say, “put him on the street.” When Petrasso finished, he faxed complimentary copies of the letter to Pons, Fleischmann, and Hawkins. Hawkins was given his copy by an acquaintance who had highlighted the lines regarding him and Hoffman. “I read it,” Hawkins recalled. “Then I got hold of Bob Hoffman, and I said, ‘Tie your boots on, buddy.’ ”’

The next day Hawkins had to answer to Pons, who was surprisingly calm and understanding and only asked him to exercise more caution should there be a next time. Hawkins believed that Pons understood the circumstances.

Maybe so. When

Hawkins

went

home,

however,

he

heard from the inimitable attorney-at-law C. Gary Triggs, who, Hawkins said, served up a different story entirely. “This Triggs character calls me up,” said Hawkins, “‘just chews me up: “You can’t say this; you can’t say that.” And I say, ‘But wait a minute, do you know what happened, and—’ ‘But you can’t say this and—’ AndIsay, ‘Look, don’t bother me.’ And hung up. Triggs was saying I shouldn’t be giving out information to people that could come back and threaten the patent position of the university and Stan and Martin and on and on.” Hoffman, on the other hand, felt betrayed by Petrasso, who only told

him at the end of the third of their four conversations that he was writing a paper. This would explain, from Petrasso’s point of view, why the fourth conversation lasted only ten minutes. “I found it very offensive, and professionally unethical,’’ Hoffman said. ““They should have said something to begin with: “Well, we want to write up an article, and we’d like to talk to you about it.’ And I would have said, Well, the conversation is at an end, you talk to Stan about what you want, because it’s his

data.”

Petrasso had no defense save extenuating circumstances. “I feel badly about it,’’ Petrasso said. ‘‘He was very honest and straightforward, more honest and straightforward than I was.”’

BAD 54

WASHINGTON,

SCIENCE

8 249

D.C.

On the night of April 24, Pons caught a flight to Washington, where he met up with Fleischmann, who was coming in from England. The two

had received an invitation from the House Science, Space, and Techno l-

ogy Committee, requesting their presence at an April 26 hearing on “recent developments in fusion energy research.”” Wayne Owens, a first-term Utah Democrat in the House of Representatives, took credit for prompting the hearing. Owens had already announced from Washington that he would draft a bill to earmark federal dollars for the establishment of a national fusion research center at the University of Utah. Certainly, the Utah delegation had been lobbying heavily in Washington to have the federal government take up cold fusion and run with it. Chase Peterson and Ian Cumming hoped to use the congressional hearings as a platform to pitch the National Cold Fusion Institute, which would be located in Utah but funded by federal dollars as well as corporate interests. The idea, as developed by Ira Magaziner, was that the university would sell shares in the cold fusion research to corporate sponsors, who would then get royalties back on the technology. The federal government would be treated just like any other corporate sponsor, which is to say the government would buy shares and collect royalties. The uniqueness of the idea may have appealed to Magaziner. All the Utah contingent needed was the right opportunity to broach this plan to the government and they would be on their way. On April 25, Pons and Fleischmann, the two “fusion pioneers,” as The Salt Lake Tribune dubbed them, were given a celebrity tour of the nation’s capital. Joined by Peterson and Jim Brophy, led by Owens, with arrangements by Cassidy & Associates, the two chemists visited virtually every influential member of Congress who had anything to say about energy policy, in particular Robert Roe, a New Jersey Democrat, who was chairman of the House Science, Space, and Technology Committee,

and Robert Walker of Pennsylvania, the ranking Republican member of the committee. The hearings, which began on the morning of the twenty-sixth, had all the makings of a cakewalk for the Utah contingent. The members of the committee seemed eager enough to write the University of Utah a blank check for cold fusion. Walker had already set aside $5 million for cold fusion out of the magnetic fusion budget, and now he suggested that

250 @ GARY

TAUBES

$25 million would be “‘the very least’? they could do. To be precise, Walker asked Fleischmann whether he thought $25 million would constitute sufficient resources to attain their goal; Fleischmann tried to divert these questions to the later testimony of Magaziner and Peterson, but

after Walker asked three times, Fleischmann sighed and speculated that

their anticipated research could cost in the neighborhood of “tens of millions of dollars.” Pons seemed nervous, as usual. Fleischmann took control, as he always

did when the two were together. Once again he appeared to be the only one in the entire delegation wary of the political risks of testifying before Congress. He later said that when he found himself sitting in the offices of Cassidy & Associates, he thought seriously about heading directly back to England. “Then I thought,” he said, “I have a responsibility; I started

this: I must continue. It’s one of those things where little by little by little you get sucked in.” In his testimony, Fleischmann had the attitude of a man who had just returned to an out-of-control situation and had to minimize the damage done in his absence. Fleischmann refused to agree, for instance, with the sociological naiveté of Dana Rohrabacher, a California Republican who insisted that much of the cold fusion criticism came from people ‘‘who are dependent on hundreds of millions of dollars’ worth of government grants, [and who] may not be open-minded.” Then he said, “You're going to put some of these people out of business, aren’t you, if you’re successful?” Fleischmann dismissed the question, but Rohrabacher refused to accept defeat. Finally Rohrabacher invoked the hallowed name of Jonas Salk, who, he said, was also not ‘“‘greeted with open arms,” yet “‘probably saved a lot of young people’s lives.” Fleischmann appeared to be wondering if this was some arcane congressional rite of initiation, in which he would have to agree to anything or lose his funding. Fleischmann even defended the aggressive response of the scientific community to cold fusion, saying it was part of the process, while insisting that neither the hot fusion budget nor any other budget should be cut to benefit his research. Fleischmann and Pons said nothing in Congress that they hadn’t said before, but they were now joined by Bob Huggins of Stanford—Huggins-Stanford, as he called himself—who gave cold fusion nearly Ivycovered respectability by testifying simply that he had done the experiment and confirmed it. He explained, however, that he would not

give details, because he was presenting his data in full at a meeting of the Materials Research Society that very evening in San Diego. Magaziner was the point man for the Utah delegation. He was there

BAD

SCIENCE

@ 251

to present his “highfalutin strategy,” much of which consisted of listing the litany of American technological inventions that had served to make the Japanese the most powerful economic nation in the world. Magaziner vividly described the Japanese “fusion fever,” then told, at Peterson’s prodding, the following horror story: ‘“When I first knew I would be testifying today, I wanted to try to get some detail on what was going on in Japan in this area. I phoned a colleague over there at . . . about 11:00 P.M. at his home. I asked him to try to make some inquiries to some friends of his in corporate research and development activities in Japan. He found them in the laboratory . . . also at work at 11:00 at night, working on the plan for this, that they’re going to develop. .. . So when I say there’s some urgency, that’s what drives me to say that.”” Magaziner suggested that it was time to take a gamble on cold fusion, “for the sake of my children and all of America’s next generation.” Chase Peterson then took the floor to pitch Utah as the place to make cold fusion happen. Among other advantages, he even suggested cryptically that Utah has “‘uninhabitable, remote regions only twenty-five minutes from the university, and twenty-five minutes from an international airport, which would serve as a useful place for what you might call special experiments.” Cold nuclear fusion weapons testing? Peterson was not saying. Peterson did not solicit money directly. He waited to be asked and then responded that, yes, $25 million might be sufficient for the time being. “Maybe that needs to be $125 million someday,” he added, “but that’s not of any importance right now.” The hearing was

divided into two

acts, the second of which was

devoted to the testimony of those who advocated hesitating before shoveling research money on cold fusion. Steve Jones gave his personal history of muon-catalyzed fusion and piezonuclear fusion, then explained the various reasons why Pons and Fleischmann’s heat was not enough to prove fusion. “Now,” said Jones, who was revealing an interesting touch for metaphors, “‘this is a tender shoot, as you can tell. It is difficult to say what it will become. Some think and suggest strongly that this is a tree, and it will grow up very quickly and provide us enough wood for all our energy needs for generations. I do not think it is. Let’s give it a chance to grow. I think adding too much fertilizer at this stage will be detrimental.”’ The last scientists to testify were Harold Furth, who ran Princeton’s plasma fusion program, and Ron Ballinger, from the MIT program. Furth suggested that if a definitive answer on cold fusion could be had within a month or two, then it could not possibly be harmful to wait and

252 B GARY

TAUBES

see. He also noted that it might not be a bad idea for Pons and Fleischmann to do those control experiments. He said if he were Sherlock Holmes, he would call this “The Case of the Missing Control Experiment.” And he would have to ponder what this missing experiment signified: ‘‘I have a feeling it does mean something, and it is conceivable to me that if this committee were to encourage Dr. Fleischmann and Dr. Pons to say something further on this topic, they may, indeed, have further things to say.” The committee, however, did not press the point. By that time, seven hours into the hearings, Pons and Fleischmann had left and the fortyeight members of the committee had dwindled to less than half a dozen. By the time Ballinger testified, also strongly advising caution, only two committee members were still in attendance. After the hearings, Wayne Owens told the hometown papers that the committee had been favorably impressed. He called the Utah delegation’s testimony a “home run.” As the Tribune headline would put it, OWENS: U. HIT HOMER OFF FUSION PITCH. Stephen Studdert, a White House staffer who hailed from Utah, then announced that Pons and

Fleischmann would be flying back to Washington in a week to meet with Chief of Staff John Sununu, who was anxious to talk science with the fusion pioneers. None

of this optimism, however,

took into account

the fact that

Peterson and company had initiated a public relations backlash with their testimony. The hearings turned disinterested parties into interested parties, although the transformation would take a week to become visible. Robert Park of the American Physical Society, for instance, in his news-

letter, would refer to Magaziner’s “‘highfalutin’’ thesis as the Pennsylvania lottery principle: “If the pot is big enough, we shouldn’t pay too much attention to the odds.’’ Robert Rosenzweig, president of the Association of American Universities, would call the Utah behavior in

Washington “‘absolutely outrageous” and ‘‘fundamentally offensive.” Even Chase Peterson came to realize that appearing in Congress with both a Washington lobbying firm and Ira Magaziner in attendance was ‘dumb politics.” Peterson later observed that they should have anticipated the reaction. “It was a whopping mistake,” he said.

BAD 55 SAN

WASHINGTON, DIEGO,

SCIENCE

#§ 253

D.C.

CALIFORNIA

Huggins-Stanford had been mysteriously evasive in the congressional hearings, although whether the representatives found his testimony suspicious was unclear. Marilyn Lloyd, for instance, who was chairman of the Subcommittee on Energy Research and Development, asked Huggins if his work has been subjected to outside reviews. “We have submitted a full professional paper to an internationally recognized journal,’ Huggins responded. Lloyd then asked, “Could we have a copy of this for the record, your Paper or statement?” Huggins replied, “That is not written at the present time.” So what was one to make of Huggins? The only unambiguous message in his testimony had been that his work had been limited by “zero funding” and that the distinguished members of the committee were in a perfect position to rectify that matter. From Washington, Huggins flew directly to San Diego for the special cold fusion session of the Materials Research Society. The MRS, like the American Chemical Society two weeks earlier, had scheduled a cold fusion session at their spring meeting to deal with the new research. Huggins would be the only one presenting positive results. And, as Nate Lewis noted later, “‘he got reamed, but, of course, Congress didn’t hear

anything about that.”’ Before Huggins’s presentation, Howard Birnbaum spoke. Birnbaum ran the Materials Research Laboratory at the University of Illinois and had worked with hydrogen in metals for much of his life. This may have prompted him to do something that neither Huggins nor Pons and Fleischmann, apparently, had bothered with. He reviewed the extensive scientific literature to see what it had to say on cold fusion. As it happened, the hydrogen-palladium system had been studied for a hundred years, which made it one of the best known of the hydride systems. All those cold fusion proponents who were claiming that something magical happens within the lattice of the palladium were strenuously neglecting the fact that what happens in the lattice was already well understood. Birnbaum reported, for instance, that the literature contained details

on the behavior of deuterium inside the lattice structure of the palladium: the ions sit randomly in the interstitial sites of the palladium and repel one

254 0 GARY

TAUBES

another, resulting in their dispersion throughout the lattice. This would make it even less likely that they would fuse. The literature even documented how far apart the deuterium ions sit within the lattice: 2.75 angstroms if they sit within the octahedral or tetrahedral sites of the lattice, 1.70 angstroms if they are forced into both. These distances, however infinitesimal, are still considerably larger than

the distance between two deuterium atoms in a deuterium molecule— say in deuterium gas, which is 0.74 angstroms, or in heavy water, which

is 1.50 angstroms. Birnbaum also explained how Pons and Fleischmann apparently had

arrived at their belief that the equivalent pressure exerted on the deuterium in their sophisticated test tubes was 10?” atmospheres. Pons had testified to this enormous number in the congressional hearings earlier that day. He called 10?’ atmospheres an ‘astronomical pressure,’ which

it certainly is. Not a single member of the House Science, Space, and Technology Committee nor the assembled staff remarked that this seemed patently impossible. (George Chapline, a Livermore physicist, later called this the most bizarre moment in the entire cold fusion affair. He suggested that to testify to a fact that happened to be wrong by twenty-three orders of magnitude “must break a record of some sort.” Even for Congress.) As Birnbaum explained it, the two chemists had apparently derived this number by misapplying a formula for figuring a quantity called the fugacity of a gas. For starters, fugacity is a parameter that is related to the energy of the gas and happens to be equivalent to the actual pressure of the gas only at very low pressures. (“‘I find it inconceivable,” Birnbaum would later say, ‘that a good chemist doesn’t know the difference between fugacity and pressure.) At higher pressures for a gas such as deuterium, the fugacity greatly exceeds the actual pressure. When Pons and Fleischmann worked through the formula and ended up with a number like 10?” atmospheres, they might have realized, had they been skeptical, that the formula was not applicable. But Pons and Fleischmann had chosen to accept the number as valid. As it happens, Birnbaum explained, deuterium can be loaded into a palladium lattice with equal efficiency by placing a piece of palladium in a canister, then pumping deuterium gas into the canister. With this method, the palladium becomes saturated with deuterium at about 15,000 atmospheres, which is the correct answer to the question that Pons and Fleischmann had attempted to calculate mathematically rather than empirically. Fifteen thousand atmospheres, however, is not nearly enough pressure to provoke nuclear fusion.

BAD

SCIENCE

§ 255

Huggins followed Birnbaum. Pons and Fleischmann, for all the attention, had been let off easy in public debates on their work. Huggins was not so fortunate. Scientists assume that when their colleagues present data they can defend it strongly and accurately. Thus, scientists receiving the information will often make extremely aggressive attacks on the evidence to test it for weak points. Science is, metaphorically, a court of law, in which one is expected to prove one’s case beyond reasonable doubt. In good science, it is considered as reprehensible to declare a wrong result carelessly as it is, in a court of law, to convict an innocent man.*” For the

first time, Huggins publicly displayed his ignorance of the details of his own experiment, and his audience, less gullible than a quorum of congressional representatives, took him to task for it. Huggins began by stating unambiguously that he would like to abolish the term cold fusion. “‘If it exists,” he said, “‘this is a solid state phenome-

non; it should be called solid state fusion.”” He explained that the colder the cold fusion, the harder it would be to extract useful energy from it. Solid state fusion sounded a lot like solid state ionics, the field in which

Huggins had made his name. (For several weeks, members of the Utah contingent used this terminology, apparently believing that it was more serious and scientific and would make cold fusion more worthy of government funds.) This may have been Huggins’s only coherent statement in his twenty-minute presentation and the hour of questions that followed.® He then described his experiment and his results, which were that a heavy water cell ran hotter than a light water cell at constant voltage. His audience suggested that quite a few chemical effects existed—the difference in conductivity between his two electrolytes, in particular—to account for this result. When asked whether he had considered these effects, Huggins responded that this effect or that “‘was small potatoes”’ compared with the heat they had seen. His transparencies, however, showed that for a particular input power, the heavy water cell ran only one degree centigrade hotter than the light water cell, a difference that was easily explicable by the difference in conductivity. But Huggins didn’t seem interested in talking about the actual data. “Small potatoes,” he said, ‘‘small numbers,’’ waving off the

questions and smiling, as though such criticisms were inconsequential. He also suggested that others had failed where he had succeeded because they had allowed ordinary water to contaminate their heavy water, thus making the effect vanish. This ran counter to quantum theory, as Steve Koonin had shown, which predicts that a mixture of heavy and light

256 & GARY

TAUBES

not seem water would increase the fusion rate. Huggins, however, did

interested in theory either. Perhaps the most telling account of Huggins’s performance was written a few days later by a young Utah chemist who described it in a note to John Gladysz at Caltech: It’s looking pretty ugly, isn’t it? I was at the San Diego MRS meeting; it was similarly bloody. I felt like poop. People there were ridiculing the P + F work. Huggins stood up to present some evidence in support of Stan’s work. Lousy presentation, highly questionable results. Add to that he hadn’t a clue of what his students did in the lab, and he came

off

looking like a buffoon. This from a guy who went into that meeting thinking that Huggins was one of the gods ofsolid state electrochem. Oh to be young and naive.*

56

PASADENA. SANTA

COLLEGE

STATION.

BARBARA

The congressional hearing on cold fusion was broadcast late in the evening on C-SPAN, and Nate Lewis watched it with growing disbelief. “I kept waking up,” Lewis said. “I couldn’t sleep; I had to watch what they were saying.” After watching the entire hearing, Lewis drove to his lab. He didn’t seem to care that it was now the middle of the night. By the time he arrived, he was irate. He cornered chemists who were working very late and launched into a tirade. Mike Heben, one of his graduate students,

recalled, ‘‘Nate really went off the deep end when he found out they were asking Congress for $25 million. That’s when he got more involved as a citizen rather than as ascientist.” The next day Citizen Lewis decided he would go public at the upcoming American Physical Society meeting in Baltimore. Lewis had been invited to discuss his cold fusion work, thanks to Steve Koonin, but

initially declined, not wanting to present his results until they were ironclad. After seeing the Utah contingent’s Washington performance, he changed his mind. The congressional testimony had served two purposes for Lewis. It had provoked his anger, and it had confirmed something that he had

BAD

SCIENCE

8 257

suspected for several weeks: Pons and Fleischmann never did the experiments in which they reported getting four times as much energy out of a cell as they put in—or eight times as much, or whatever. This was the famous third column in the table in their paper: excess heating as a percentage of break even, with all those remarkable numbers: 286 percent, 1224 percent, 438 percent, and 839 percent. All were well above break even. What did this mean? Lewis later said he finally figured it out when he was playing basketball and thinking about the third column. To run a cold fusion cell, he explained, you had to pay two energy prices. The energy to run the cell had to overcome the electrical and chemical resistance of the cell, and it had to reduce D,0O into its compo-

nent parts: deuterium and oxygen. What the third column suggested was a cell in which extra deuterium gas bubbling off the cathode was simultaneously pumped back around and down on the anode. Thus the cell could be run without ever spending energy to reduce the D,O and make oxygen. In chemical terms, Pons and Fleischmann had depolarized the anode. It seemed simple and clever, and it made sense. It meant Pons and Fleischmann must have done yet another set of experiments that had not been discussed openly. In fact, Pons and Fleischmann said in the footnote that the cell potential was only 0.5 volts,”° while they had said in the body of the paper that it required 0.8 volts to charge the cell. So they must have saved that 0.3 volt difference by depolarizing the anode. The problem, however, appeared to be that it actually took 3 to 4 volts and sometimes as much as 10 to force current through the cell. To achieve their remarkable numbers, Pons and Fleischmann had somehow managed to reduce 3 to 10 volts down to 0.5 volts. Lewis had explained this revelation to his researchers and had them try the experiment. They failed. They couldn’t generate enough deuterium gas to depolarize the anode. So Lewis called Chuck Martin at Texas A&M and told him what they’d done. Martin tried it, and he failed. Meanwhile, Lewis tried the brute force approach: ““We just got some deuterium gas—it’s very expensive, but we said, ‘Okay, we’ve blown enough money, we might as well continue’-—so we opened it up and bubbled it through and we could never get the voltage down.” Lewis called Martin back and suggested Martin try to depolarize the anode by bubbling deuterium “through there like crazy.’’ He added that “maybe we’re screwing up.” Martin said he couldn’t do it either. At that point Lewis suggested that it couldn’t be done. He strongly urged Martin to ask his friend Pons

258 #8 GARY

TAUBES

about the experiment. Lewis said, “So all right, Chuck, I know you're talking to Stan. Ask Stan if he ever really did this experiment, or if the four-to-one and ten-to-one numbers, the numbers in the newspapers, the only numbers they’ve ever reported, are ones he actually measured, or calculated using some assumption.” “No,” Martin said, ‘I can’t do that.” “Chuck, we can’t do this experiment,’’ Lewis said. “‘It can’t be done

unless we’re really missing something. You’ve got to ask him.” So Martin phoned Pons. The conversation went like this: “We measured everything,” Pons said. ““There’s nothing wrong with this.”’ “But Nate says,” Martin replied, ‘that you can’t do this.” ‘What do you mean we can’t do this?” asked Pons. “Well,” Martin said, “you can’t do it. We can’t do it. Nate can’t do

it. We don’t think it can be done.” “Oh yeah,” Pons replied. ““There were some assumptions somewhere, but we measured every other number.” “Okay, which numbers did you measure?” Martin asked. Pons wouldn’t tell him. As Lewis put it, “Chuck kind of reached what Stan wanted to hear. So Chuck called me back and said, ‘I don’t think he measured them either. He wouldn’t tell me for sure, and I think I pushed himalittle bit too far, but I don’t think he measured those numbers.”

What Lewis heard in the congressional testimony was Fleischmann confirming his hunch that those numbers had never been measured. Fleischmann admitted that his figures were speculative, but he justified the speculation, or projections, as he called them, as “part and parcel” of fusion research.”! He reiterated several times that this was speculation, although he never phrased it in such a way that a casual observer or a congressional representative would realize exactly what he was talking about. “This would be a much more energy-efficient device,” Fleischmann said, ‘‘and underlies one of the results which I'll show you, one of the sets of calculations . . . which is really a hypothetical energy release— I’d stress that—because it involves the recalculation of our data to project them to the condition. . . .”” So Pons and Fleischmann had only speculated that they could generate more energy ina fusion cell than they consumed. They would have to build a perfect cell, which Fleischmann called a “more sensible” way to do it, or a more “energy-efficient” cell, or a projection “‘to a viable technology.” Indeed, Lewis later observed that it would have been more sensible, and more energy efficient, had it been thermodynamically pos-

BAD

SCIENCE

§ 259

sible, which it wasn’t. It was just another way of trying to get a free lunch out of thermodynamics, which never gives free lunches. The bottom line was that cold fusion had captured the imagination of the world on the basis of mere speculation, or gross speculation, depending on one’s point of view. Until April 26, no one but Pons and Fleischmann knew this for certain.”° On the morning of April 27, Lewis called Steve Koonin and explained the origin of these numbers. Koonin was outraged. Koonin had planned to speak at the APS meeting; now he decided he was “going to come out really strong and say this is all nonsense.” He called up the lawyer at Caltech and asked him how far he could go in public without being vulnerable to a lawsuit. As Koonin tells it, “He said, “Well, you can’t use the word fraud, and you can’t imply any deception.’ So I thought long and hard about exactly what I was going to say.” A few days later, he rehearsed the talk at the annual spring meeting of JASON. He asked Dick Garwin, Will Happer, and others who had more experience explaining science in public for advice. Richard Muller, a physicist at UC Berkeley, gave him what he considered the best advice: “that is,” said Koonin, “leave no doubt about where you stand, because

if you leave at all the possibility that this could be right, it will be misinterpreted by the press. I figured, what the heck, I have tenure. I believe 99.99 percent that this is nonsense at this point. So I just said, “Okay, I’m going to shoot the moon.’ ”’ Meanwhile, Nate Lewis had contacted the APS organizers and been added to the schedule. With Charlie Barnes, who was overseeing the

physics side of the Caltech cold fusion effort, Lewis went to see Barclay Kamb, the Caltech provost. The two scientists briefed Kamb on the situation as they saw it and what they would like to say in Baltimore. Whatever instructions they received, the way Lewis and his chemists told it later was that the provost said, ‘‘Say anything you want, just don’t use the F-word.”’ The F-word, of course, was fraud.

260 8 GARY

TAUBES 57

SALTLAKE

CITY.

GROUND

ZERO

Peter Dehlinger saw Stan Pons for the last time on April 29, which seems to have been a frenetic day even by the standards of the previous month. The new patent lawyers from Houston were in town, and Dehlinger had flown in to pass the torch. C. Gary Triggs, Pons’s personal lawyer, was also in attendance. They all met Chase Peterson at his house along with Ian Cumming, Ira Magaziner, and Richard Giauque, the Salt Lake

City lawyer. And on campus Pons and Fleischmann had another meeting, this one with John Bockris and Supramaniam Srinivasan, two Texas A&M chemists who had arrived to talk cold fusion. Pons appeared totally besieged. “‘He was extremely impatient,” Dehlinger recalled. ‘Extremely. I don’t think I’ve ever seen a man driven to such impatience, almost as though a stone in the shoe would have driven him wild at that point . . . the way people behave when they’re under far more pressure than the good Lord meant them to be under.”’ Fleischmann was much more in control, which Bockris noticed as well. Several times over the course of the day, Bockris asked Hugo Rossi to watch out for Pons. He was concerned, said Rossi, ““with Pons’s mental health.”’ As Bockris described it, Fleischmann ran all the interactions while Pons cowered in the background muttering of his critics, “Bastards, bastards,

all of them.” Bockris was Fleischmann’s senior by two years and had taught the fusion pioneer at Imperial College back in the 1940s. He was not, however, the most credible of sources.

Into this pressure cooker came Rulon Linford to see about cementing the collaboration between Los Alamos and Utah. In his congressional testimony, Pons had promised the committee members several times that the Los Alamos collaboration was imminent. ““We have nineteen new experiments being set up now,” Pons had said. “One of those is a demonstration of a previously run experiment for Los Alamos. They will come up, make the measurements they want to make on our own system, bring their electrochemists, and let the electrochemists go through our method of measuring the thermal output, and when they are satisfied with what they see, then they will take that experiment away.” Los Alamos was ready to go. A team of chemists and physicists would fly up to Utah the moment Pons told them directly that he hada cell generating excess heat. The laboratory’s lawyers had already drafted a five-page collaboration agreement. The collaboration, however, was undone by Pons’s feelings of perse-

BAD

SCIENCE

# 261

cution and then by the local lawyers. First, Linford had a run-in with Pons, sparked bya slight Pons felt he had suffered at the congressional hearings. Harold Furth of Princeton had called Linford before the hearings to learn exactly what Pons had said about his light water controls in his Los Alamos seminar. Linford, who had a videotape of the seminar,

found the point at which Pons answered the question about light water—that he had seen heat and then discontinued the experiment— and played it for Furth over the telephone. In Washington, Furth had apparently confronted Pons with what he had said in Los Alamos, suggesting it was proof cold fusion did not exist. Pons had not taken it well. Now Linford stopped by the Utah lab as Pons and Fleischmann were showing the Texas A&M people around; then he slipped away for a few minutes with Pons. As Linford recalled, Pons pointed to a bubbling cell and remarked that this was the one he expected the Los Alamos people to retrieve and analyze. Then he got down to what was bothering him. He wanted to know if Linford was a friend or not. “T asked him what he meant,”’ said Linford, ‘‘and he said he had been

somehow cornered by Harold [Furth] in this heavy water—light water stuff. I shrugged my shoulders and said, ‘When you came down and gave the talk, we told you we were going to tape this. And when he called me up and said was this on tape, I said yes. And I don’t consider that friend or foe, just the facts.’ And so I think he viewed me with suspicion,

because I had and do have all these ties with all the magnetic fusion people around the country. . . . He was suspicious about that.” As Peter Dehlinger observed, ‘“‘Pons hadareal distrust of the people at Los Alamos. He was sure they were gunning for him.” It may not have mattered, because if the local powers-that-be had wanted a collaboration, they might have forced the issue. The dilemma at Utah was how to exclude the federal government from taking over cold fusion while simultaneously convincing Congress to invest in a cold fusion institute, which now seemed like a very real possibility. On the one hand, Norm Brown later said that he suggested that if they went ahead with the Los Alamos collaboration, they would be in essence giving the technology to the Defense Department, and federal law was flexible enough to accommodate the government’s ownership of cold fusion if such should be the government’s desire. Linford, on the other hand, believed that the Los Alamos lawyers had drafted a contract that made it very plain that the government had no rights whatsoever to the technology, but he wasn’t making the decision. By now, Ian Cumming and Jon Huntsman, another local businessman

262 8 GARY

TAUBES

and university benefactor, were having serious discussions with General Electric about buying into the research. Said Hugo Rossi, “General Electric advised very strongly that we should not get involved with the DOE. Of course, it was self-serving. They felt that it should be developed in the corporate setting, not in the national laboratory setting.” If the Utah contingent could get GE to commit, so the logic went, then they wouldn’t need Los Alamos. It was all touch and go. Which partner would take them to the dance? When Linford returned to Los Alamos, Jim Brophy told him that he seemed to have eight lawyers working on cold fusion now, and they were all ‘‘on his back.” In other words, the

Los Alamos—Utah collaboration was going nowhere.

58

THE

PRESS

On April 30 The New York Times, which traditionally serves as the paper of record for the scientific community, declared cold fusion dead. According to the Times the University of Utah could “‘now claim credit for the artificial-heart horror show and the cold fusion circus, two mile-

stones at least in the history of entertainment if not of science.”’ As for the $25 million that Chase Peterson had requested in Congress, said the Times, offering up another variation on Pascal’s wager, “the Government would do better to put the money on a horse.”’ The next day the Boston Herald denounced cold fusion in a page-one story. Ron Parker of MIT was a principal source and quickly came to regret getting involved. He thought he could manipulate the press, which made the episode something of a morality tale of science and journalism. The story, which Parker later called “‘a real zinger,”’ ran under the kind of one-inch banner headlines usually reserved for the end of wars or the resignation of presidents: MIT BOMBSHELL KNOCKS FUSION “BREAKTHROUGH” COLD

Leading scientists at MIT derided a breakthrough cold fusion experiment as “‘scientific schlock,’’ and will release a report today charging two University of Utah chemists with misrepresentation and ‘’maybe fraud.’’

BAD

SCIENCE

@ 263

Parker was quoted saying, “Everything I’ve been able to track down has been bogus, and I think we owe it to the community of scientists to begin to smoke these guys out.” Motivating all this was Parker’s new-found anger. He had become increasingly confounded by the congressional testimony and the local and national reporting. Ron Ballinger had returned from Washington and given Parker a play-by-play of the Utah delegation’s performance. Parker later described it as ‘‘asking for $25 million. You know . . . send us a check, or, preferably, we have briefcases, fill em up with cash.” And

he was still outraged by The Wall Street Journal Pons-cum-Rutherford editorial that had accused MIT physicists of being petty and smallminded. “‘It besmirched the memory of Rutherford,” Parker said, “‘“who

happens to be one of my heroes, and insulted my institution.” And finally Parker was angry at The Boston Globe, which he felt had been cheerleading for cold fusion rather than reporting the facts. Parker had stopped talking to the Globe, but he continued to talk to reporters from The New York Times and The Washington Post, who he felt were keeping their perspective. When the Globe reporter, Richard Saltus, wrote Paul Gray, the president of MIT, complaining about Parker’s lack of cooperation, Gray sent the letter on to Parker. The Globe’s complaint fueled Parker’s annoyance. He decided to spite the paper and give an interview to the Herald, a Rupert Murdoch—owned tabloid that was notorious for its sensational news reporting. “Normally,” Parker explained, “I would certainly not talk to the Boston Herald, but on this particular day, I was pretty aggravated at the Globe and | thought, Well, I'll not only talk to the Herald, but I’ll really lay it on the line. So, Ron Ballinger and I had the Herald reporter over here [on Friday] for a good hour and a half. And, we really lit in to Pons and Fleischmann. I had become convinced through our own work here that they had—to be charitable—misinterpreted what they were seeing.”’ Parker asked the reporter to hold the story until Monday, May 1, the day of the APS meeting. Late Sunday night, Parker received a call from CBS Radio. “Dr. Parker,” the reporter said, “we see that you’ve accused

the Utah people of fraud and scientific schlock. We’d like to know if you'd elaborate on this.” To which Parker replied, ““What?” Parker later said that he had made the fraud and schlock comment in reference to Pons and Fleischmann’s manipulation of the gamma ray spectrum. In a more conventional scientific episode, this might certainly have been described as misrepresentation at the least. In cold print,

264 I GARY

TAUBES

however, out of context, without sufficient explanation, his words went

much further than Parker wanted to go or thought he had gone. Monday morning Parker contacted the MIT public relations office and scheduled a press conference. He intended publicly to deny the story, or at least the quotes, which he was finding increasingly difficult to believe he’d actually said. The press officer, Eugene Mallove, then took the opportunity to chastise Parker: ““You cannot win by playing with the press,” he said. ‘“You’re always going to lose.’’”* Parker tried to deny the quotes, but Nick Tate, the Herald reporter, pointed out that he still had the tape of the Parker-Ballinger interview. They listened to it, and each thought it vindicated his position. The following day, the Herald ran the transcript of the interview. And,

indeed, Tate had quoted Parker correctly, although Parker still argued that Tate had asked leading questions and quoted him mildly out of context. “I think we both lost,” Parker said. ‘And I think I got tarnished in a way, because people thought, Jesus, you know, Parker’s another raving maniac at MIT.”

59

BALTIMORE,

MARYLAND

Neither Stan Pons nor Martin Fleischmann appeared at the American Physical Society meeting in Baltimore on the evening of May 1, although both had been invited. The organizers had hoped to get Fleischmann because they had been told that “he would speak to physicists better,” but they got a message that neither of the chemists could make it because they were too busy preparing for their congressional visit. This made the fusion pioneers appear more interested in getting money from Congress than in being vindicated by their peers. But Pons, at least, knew in advance what would happen at the APS meeting; he would be crucified. In fact, the session that took place may have been even more brutal because the main targets did not appear. The first speakers were Steve Jones and Jan Rafelski, who, in this crowd, passed for the pro—cold fusion contingent. Jones, wearing a sedate dark suit, played down his data, although he did showaslide of his APS abstract, which he had submitted back on February 3. Showing the abstract constituted staking his claim to priority; he called it ‘“‘a modest

BAD

SCIENCE

8 265

submission” and laughed discreetly. Then Jones went about distinguishing his results from Pons and Fleischmann’s, which meant establishing that he had neutrons and they had heat alone, and that to claim fusion on the basis of heat alone was irresponsible. ‘‘The ratio of our results,” he said, ‘“‘to the Pons-Fleischmann

results is about the same

ratio as

between the dollar bill and the national debt.” Steve Koonin and Nate Lewis followed and nearly killed cold fusion between them. Koonin later said that someone told him that he hit a triple and Lewis hit a home run. “He was good,” Koonin said about Lewis. “People were just stunned.” Koonin described the problems with the theory of cold fusion, and the forty to fifty orders of magnitude that had to be bridged; then he discussed and rejected all the possible bridges. He said the BYU result was theoretically dubious but not impossible. ‘The situation was gloomy but not yet terminal.” “My conclusion,” Koonin said, “‘based on my experience, my knowledge of nuclear physics, and my intuition, is that the experiments are just wrong. And that we’re suffering from the incompetence and perhaps delusion of Drs. Pons and Fleischmann.”’75 Lewis then articulated the view from the experimental side of the affair. He had a confident, staccato style that further pointed up his knowledge of the material. He described the results of the extensive Caltech cold fusion studies, which could be summarized as no neutrons, no gamma rays, no tritium, no helium, no excess heat. No nothing.

“Heat was confusing to nearly everyone,” he said, “including all my electrochemical friends, and I will try to walk you through this and you will be shocked. Guaranteed.”’ Then Lewis explained how the calorimetry was done in his lab and in Pons and Fleischmann’s, and how they derived their numbers, and all the assumptions they made. He noted that the measured power of Pons and Fleischmann’s own cells was always less than the input power, which is to say, the ‘‘cells were not good fusion heaters, they were great fusion refrigerators.” Lewis cited a 1958 paper by Abner Brenner of the National Bureau of Standards that documented the result of electrolysis of water in a cell with a platinum anode and palladium cathode. “‘Not fusing,’’ Lewis said, but other than that it sounded very familiar. Brenner found that if the cells—the cold not-fusing cellsk—were not vigorously stirred, they would generate definitive temperature gradients, which is to say, they would be hotter in one place than in another. One could then observe apparent excess heat by the placement of the thermometer. Pons and Fleischmann had not stirred their cells, and Lewis came down hard on this point.’

266 § GARY

TAUBES

“The errors,” he said, ‘‘are obviously going to be very large.’’ Then Lewis described how the third-column numbers originated and the attempt to circumvent the laws of thermodynamics. Lewis concluded that if Pons and Fleischmann, among others, were

going to have publication by press conference, he would institute peer review by press conference. He presented a lengthy list of pertinent questions that reporters might want to ask Pons or Fleischmann, should they get the chance. ‘‘That’s what I have to say,” Lewis said. ““I hope I’ve clarified some points.”’ The audience of 2,000 physicists gave him a standing ovation. Lewis later remarked that it made him think of retiring. “I felt like I was Kareem [Abdul-Jabbar, of the Los Angeles Lakers] or something. They could sit me down. Give me a Rolls-Royce. I’d never been through anything like that in chemistry before.” The Utah press, and Pons, Fleischmann, and the proponents of cold

fusion, however, later portrayed him as a petty and small-minded soul who was worried about losing his funding. He became another de facto fusion physicist. Lewis received hate mail from Utah. He refused to reveal what the letters said, but noted that he gave them to the police.

He later rejected an offer by the University of Utah physics department to lecture, because the letters had been “‘too weird’’ and he was not anxious to visit the state. Lewis countered that if cold fusion were true,

electrochemists would all have funding beyond their wildest imaginations. “‘An electrochemist’s wet dream,” he said.

After Lewis a half a dozen other physicists reported negative results, including Stan Luckhardt from MIT, and Moshe Gai, a Yale physicist whose results Nature would announce on its cover two months later. Robert Park in What’s New described the meeting as an unrelenting assault on cold fusion. Nonetheless, he predicted that “‘the corpse of cold fusion will probably continue to twitch for a while,” which turned out to be prescient but understated. Of all the scientists present, it was Jan Rafelski who made the most revealing comments on the state of cold fusion. Rafelski observed that it was “totally irresponsible” to claim nuclear fusion on the strength of excess heat results. Thus, the Pons-Fleischmann claim was meaningless. He then admitted that a miracle was needed to explain even the infinitesimal effect described in the BYU paper, of which he was a coauthor. Still, he said, “the intensity of current interest (many serious

efforts to duplicate both experiments) shows that the possibility is there for a great future discovery. So we really are out to seek the miracles.” The logic was relentlessly circular.

BAD

60

WASHINGTON,

SCIENCE

§ 267

D.C.

After the APS special session, four of the society’s officers went to Washington to meet with various congressional representatives and committee staffers about funding for physics. Steve Koonin, who was chairman of the nuclear physics division of the APS, was accompanied by Robert Eisenstein of the University of Illinois, Virginia Brown of Livermore, and James Ball of Oak Ridge. The four physicists were scheduled to see Robert Walker, the ranking minority member of the House Science, Space, and Technology Committee. Walker had been the most forthcoming about donating millions to Utah’s cold fusion future, and the physicists spent a long time trying to convince him that this was not the sensible thing to do. “Isn’t there a chance that this could be right?” Walker asked. “Shouldn’t we be putting in just five million dollars, if we put all this money into magnetic fusion? Just suppose that they’re right.” “If somebody walked in here and said that they’d invented an antigravity machine,” Koonin asked, ‘“‘would you give them money?” “Yes,”’ Walker said, “I probably would.”’

61

SALTLAKE

WASHINGTON,

CITY. D.C.

Pons and Fleischmann were scheduled to meet with John Sununu in Washington on May 4. That morning they were told that Sununu had canceled because of “‘scheduling conflicts.” On the same day, the House Science, Space, and Technology Committee announced that it was canceling afield trip to Salt Lake that had been scheduled for May 13. Jake Garn, one of Utah’s senators, had planned to fly a dozen committee members to meet with the fusion pioneers and witness a private demonstration of cold fusion. Congressman Wayne Owens, who was spearheading the cold fusion effort in Washington, then told reporters that the recent criticisms had

“‘caused the bloom to go off the fusion rose.”’ He added that he, at least, was not giving up; Cassidy & Associates would still draft legislation to

268 #8 GARY

TAUBES

erect a cold fusion institute in Utah with the help of $50 million in federal allocations. If necessary, he said, he would gladly settle for $25 million.

62

THE

PRESS

While Pons and Fleischmann were being stood up in Washington, the May 8 issues of Time, Newsweek, and Business Week appeared on the newsstands, with cold fusion on the covers of all three. The magazines had gone to press before the American Physical Society meeting, so the writers were taking their chances. The articles were pedagogical examples of what A. J. Liebling, The New Yorker's legendary press critic, would have called on-the-one-hand-this-on-the-other-hand-that journalism. He would have needed three hands, however, for cold fusion.

Time (FUSION OR ILLUSION? HOW TWO OBSCURE CHEMISTS STIRRED EXCITEMENT—AND OUTRAGE—IN THE SCIENTIFIC WORLD) took the most critical view, the rationalist’s view, assuming that even if cold fusion were

right, it was probably no panacea, and then again it probably wasn’t nght. (The Time correspondents, at least, had called around the cold fusion

community to get an advance line on what might be said in Baltimore.) As the article put it, “The solution to the world’s energy crisis is not likely to be declared in a press conference. It must be slowly and carefully worked out, step by painstaking step.”

Business Week (FUSION IN A BOTTLE, MIRACLE OR MISTAKE, A SCIENTIFIC DETECTIVE STORY) took the middle of the road: ““The hope that cold fusion is the answer

is understandable,”

the article ended,

“but the

doubts about it are reasonable. The scientific jury is still out.” Newsweek (THE RACE FOR FUSION, THE SCIENTIFIC DEBATE—WHY THE STAKES ARE SO HIGH) took the populist angle. Although the magazine suggested several times that Pons and Fleischmann could be wrong, it treated the possibility as trivial in light of the excitement provoked. On the one hand, Newsweek said, ““The frenzy has distorted the usual careful

pace of research, producing sloppy science. Results that turn out to be wrong are sending other labs down blind alleys, and the carnival atmosphere may also raise expectations for fusion that can’t be met overnight.”” And on the other, the article concluded, ‘‘Pons and Fleischmann

BAD

SCIENCE

8 269

have unleashed a revolution in the way scientists think. That’s not as significant as lighting the world, but in a discipline that depends on inspiration and imagination, that’s cause for celebration.”’ It seemed like a journalistic variation on Jan Rafelski’s belief that the intensity of interest, by definition, makes an activity worthwhile. There

must be a miracle, or else why were all those scientists doing it? On the other hand...

63 LOS

SALT

LAKE

ANGELES,

CITY.

CALIFORNIA

Curiously enough, Stan Pons and Martin Fleischmann professed to be elated by the results that had been reported in Baltimore. “We are extremely pleased because they confirm our findings,” Pons told The Deseret News. ‘“The absence of neutrons doesn’t concern us in the slightest. We couldn’t be happier. We and other scientists will soon tell them why this is so.”’ What Pons apparently had in mind was a new twist on the nuclear by-products of cold fusion. As he now seemed to believe that his cells were generating helium 4 rather than tritium or neutrons, the fact that nobody in Baltimore had reported observing neutrons from cold fusion was a good thing. The BYU work had not been confirmed. Now all that was left to prove was that the Utah cells had generated helium 4, and that it was a signature of cold nuclear fusion. Neither of these was inconsequential. Pons was planning to announce the helium 4 observation officially in Los Angeles at an Electrochemical Society meeting on May 8. The Utah administration was playing up this conference as another match between physicists and chemists. Hugo Rossi, who had just been put in charge of the Utah cold fusion research effort, told The Deseret

News that May 8 would be “‘F-day,”’ meaning apparently fusion day. He predicted that a half dozen papers would be presented “reporting results consistent”? with those of Pons and Fleischmann.” Pons armed himself for the Los Angeles meeting with the defense that would invert established scientific methods for good in cold fusion. “Criticism doesn’t bother me,” he told The Deseret News. ““What bothers

270 8 GARY

TAUBES

me is that these people are not doing the science. They are philosophizing, using a physical theory, parts of which may be totally wrong... . Pray, where are their data? We are presenting ours. Why don’t they publish their experimental procedures so we can see what they are doing to disprove our work? It is very easy to pontificate and to say someone’s data is wrong; proving it is another matter.”’ It was no longer a scientist’s responsibility to defend his research but the scientific community’s task to defend its criticism. Cold fusion existed until proven otherwise. The Electrochemical Society cold fusion symposium was the perfect forum for the kind of scientific debate this logic would foster. The ECS administrators wanted to avoid a repetition of the rampant negativity of the Baltimore American Physical Society meeting. Speakers would present ‘‘confirmation results” only. The official announcement said that Pons and Fleischmann would be featured on the program, followed by Steve Jones, and then “‘research groups who have verified the initial reports of Professors Fleischmann and Pons, or Professor Jones.”””* Robert Park, sardonic as ever, suggested that the ECS announcement was “‘a forgery by some evil prankster bent on discrediting the Electrochemical Society.”” Unfortunately, such was not the case. After Baltimore, Pons seemed most concerned that Nate Lewis not be on the schedule. Lewis, however, was at least as well connected as Pons,

and he appealed to his highly placed friends in the ECS hierarchy. The society then managed an acceptable compromise: Pons and Fleischmann would speak first; Lewis would speak last.” In deference to Pons’s wishes, the meeting would not be videotaped,

perhaps to prevent the kind of embarrassment that the Los Alamos tape had subsequently caused. Cameras and recording equipment were also banned. “‘An apparent effort to preserve deniability,”’ suggested Park. The press would be allowed to sit in on the session, but only after paying a $200 registration fee, which, in retrospect, was not the best politics.

Despite the precautions against negativity, the preliminary line on the session had it that if Pons and Fleischmann were to convince anyone of the validity of their data, as Steve Koonin put it, they would have to pull a rabbit out of their hats. And the audience seemed less than friendly, especially considering that the bulk of the 1,600 present were fellow chemists. One Brookhaven metallurgist described the crowd as “‘ornery.”” Their dispositions were aggravated further when they were delayed for an hour in the lobby of the Westin Bonaventure Hotel, under the hot glare of television lights, while hired security guards

cleared and searched the auditorium. What were the Pinkertons looking

BAD

SCIENCE

@ 271

for? Hidden recording devices? Bombs? Appearing before an audience that had been subjected to this kind of indignity, Pons and Fleischmann had an uphill battle. Hugo Rossi later called Los Angeles “‘the end of the innocence.” Pons reported one interesting piece of data: they had witnessed, he said, a two-day burst of heat amounting to a staggering 5000 percent of break even. But he offered nothing to support this claim other than his word asascientist. Fleischmann’s contribution to the existing body of knowledge was a short videotape made with the help of Canadian Broadcasting Corporation reporters. The video was intended to negate Nate Lewis’s criticism that their calorimetry was in error because they did not stir their cells. “Gas evolution,” Fleischmann said, by which he meant the bubbles of

deuterium and oxygen coming off the electrodes, “‘is the most efficient method of mixing known to man.” He added that he was going to “lay this ghost to rest.”” The video showed an effervescent cold fusion cell into which a red dye was poured. Within twenty seconds the dye had been evenly distributed throughout the flask.’ Fleischmann said this proved that ‘‘the argument of ineffective mixing doesn’t hold water.” The demonstration was impressive; however, it was bogus. Even if the cold fusion cell had huge temperature gradients—asay, fifty degrees hotter on one side than on the other—the red dye would have diffused evenly within a very short time. The temperature gradient in the flasks simply had nothing to do with what could be called the red dye gradient. This also begged the question of why Pons and Fleischmann hadn’t simply set up thermometers in a cell and collected the temperature data, which would have been less trouble and more convincing than the video. Later Dick Crooks of MIT suggested that the video might have been a product of Fleischmann’s peculiar sense of humor. “It’s very common in science,” he said. “When you don’t have the results, you start telling jokes.” Crooks had come to the conference with Stan Luckhardt to offer Pons and Fleischmann help with their helium analysis and to confront them on the gamma ray spectrum, which, thanks to Richard Petrasso, they knew to be worthless. Petrasso had spoken in Baltimore, but he had been

late in the program, and by the time he presented his analysis most of the audience and all of the press had left. Now, in the question and answer session following the presentations, Luckhardt put the question to Pons. Luckhardt said that they’d analyzed his spectrum and concluded that the 2.2 MeV peak had to be an artifact, and he explained why. “Do you have an answer to these criticisms?” Luckhardt asked.

272 0 GARY

TAUBES

“Our spectrum has been questioned,” Pons responded. “We have now installed a larger gamma ray spectrometer, and we’re collaborating with physicists. All we know is our peak appears when the cell is on and not when the cell is off.”’ “Can you explain why your peak is so narrow?” asked Luckhardt. “No,” said Pons in a listless whisper. At that point Fleischmann gently took the microphone. “I’m well aware the peak is wrong,” he said. “‘I’ve discussed this with Stan. It perturbs me. That was not meaningful using the equipment we had at the time.” The two chemists were asked about their light water experiment, which Pons had reported generating excess heat. ‘That’s complete nonsense,’’ Fleischmann said, “total nonsense.”

They were asked about Pons’s announcement in Utah that they had discovered large amounts of helium 4. “I didn’t make that statement,’’ Pons said.

Then Crooks took the microphone and said that they had recently performed helium 3 and helium 4 analyses from light water and heavy water cells. “In a cell such as yours,” Crooks said, “‘running 200 hours,

helium 4 in the lattice should be 10® helium atoms per centimeter cubed, which is 107 above our detection limits. And we detect no helium 3 or helium 4. We would like to extend our facilities for doing this very difficult analysis to you.” “We appreciate that,” Fleischmann said. “It’s certainly in our minds to have our rods analyzed.” Crooks added that they could do the analysis in three days. Fleischmann didn’t respond. A gentleman from Sandia National Laboratories in New Mexico then made the same offer. The press conference that followed was even more dismal. Pons looked increasingly morose. He was perspiring, sallow skinned, and grim. Fleischmann fielded the vast majority of the questions because Pons could not muster the energy to make himself heard, even with a microphone. The reporters seemed finally to have lost their patience; maybe they felt they’d been taken advantage of, and now they knew the right questions to ask. Pons and Fleischmann’s only defense was to say that they had never made statements that the reporters had heard personally, which is not effective diplomacy under any circumstances. Fleischmann blamed the press for the madness, which was like lighting a stick of dynamite and blaming Alfred Nobel for the ensuing explosion. Despite all his earlier protestations, Pons found himself sitting on a

BAD

SCIENCE

@ 273

podium with Nate Lewis, who during the scientific session had repeated his emphatic Baltimore presentation—no neutrons, no gamma rays, no tritium, no helium, no excess heat.

In the first question of the press conference, Lewis was asked to summarize his results so that Pons and Fleischmann could respond to them. Neither one had an answer, however. Fleischmann said only that he would have to see Lewis’s results published before he could make an assessment. Fleischmann then explained that they wouldn’t take up the offers from MIT and Sandia to have their electrodes analyzed because they were not “free agents’; they had a prior commitment, although he refused to say what this commitment was. He explained that they hadn’t reported the helium 4 data because the analysis was more difficult than they had expected. If they didn’t find helium 4, he said, it would “‘eliminate a very substantial part of our view of this process, yes.’’ What he didn’t say was that they had dramatically revised their talk just before the Los Angeles meeting and had decided, as Hugo Rossi later explained, that their helium data “would not go over at that meeting.” The assumption seems to have been that there are two levels of scientific data: one that can be defended against a roomful of reporters and one that can be defended in a scientific meeting. The helium data, obviously, were not of the second

type. At the end, Pons looked like a man attending his own wake, and

Fleischmann was left promising the reporters that if cold fusion was wrong, he would be the first to admit it. At the press conference in Los Angeles, however,

cold fusion first

demonstrated just how resilient it would be. The press may have finally awakened to the fact that Pons and Fleischmann had no meaningful data, but it would make little difference. For the first time ina scientific setting, the two were no longer defending their data, or lack thereof, alone. Steve Jones, once again, told the press of his observation of neutrons (although he reiterated that his results ““do not confirm in any way the conclusions of excess heat’”’). Bob Huggins of Stanford, and John Appleby and John Bockris of Texas A&M were also on the podium. These three must have seen their fusion futures slipping away, because now they stepped in to salvage the situation. Huggins, introducing himself as ““Huggins-Stanford University,” said that, regardless of what anyone else might have said or heard, “there is conclusive evidence that there is a large amount of heat generated here. .. . Clearly the thing is big. The thing is important in a commercial application sense.”’

274 © GARY

TAUBES

Appleby, a dapper Englishman and a chemical engineer by training, also reported observing excess heat and suggested to the press that what they had was deuterium fusing with helium 4 to give lithium 6. This fusion scenario would leave no detectable by-products save heat. “‘In which case,” Appleby said, “one would detect absolutely nothing.” This was an interesting twist: no one had found any fusion by-products other than heat, so Appleby was proposing a fusion process that wouldn’t provide any. It had the advantage of being virtually untestable. Helium 4, however, is an extremely stable atom. To fuse anything with it would be quite a trick, even by the relaxed standards of cold fusion. Even Jones said this was too exotic for him, and he was willing to believe that he had discovered an effect that was forty orders of magnitude on the exotic side. Finally there was the sixty-six-year-old John O’Mara Bockris, who appeared in an off-white suit that gave him the air of a plantation owner who had fallen through a time warp. Bockris had been a major player in electrochemistry for half a century. ““Everybody but everybody in electrochemistry has heard of Bockris,” said an Oxford-trained metallurgist from Brookhaven,

who

added, “I would think that if I was as well

established as he was, I would hesitate to dip my toe into a pond like that.” Bocknis professed to have no doubts about cold fusion and told the reporters in Los Angeles that Nate Lewis was the only scientist in the world who believed the Utah heat measurements were faulty. “I think it’s important,” said Bockris, “‘to note the hints constantly made by Dr. Lewis that there’s something faulty about all this heat. It is really he who is the only one who says this.” Bockris said he knew personally of fourteen labs, from six different countries, that had measured both heat and nuclear by-products. “So you can’t neglect that,” he said. ““There’s no use going on denying it with the rest of the world verifying it.” He added that his calorimetry at Texas A&M was not what Lewis had used at Caltech or, for that fact, Pons at Utah, so none of Lewis’s criticisms, or anyone else’s, applied to his own

observations. He refused to discuss his work until it was accepted in a scientific journal, but noted that his results ruled out a chemical reaction as the source of his excess heat.

BAD 64

COLLEGE

SCIENCE

§ 275

STATION

In fact Bockris had discussed his mysterious results in the scientific session, reporting that massive amounts of tritium had appeared in his cold fusion cells. These results were so remarkable that they seemed to be definitive proof of cold fusion, that is, if one believed them at all. After

Bockris spoke, Stephen Feldberg, a respected electrochemist at Brookhaven, remarked to him that neutrons ought to appear with the tritium, and as Bockris hadn’t mentioned any such radiation, he suggested that Bockns worry about the health of his researchers. Bockris replied that that was an interesting point, then commented enigmatically that maybe someone had “‘spiked”’ his cold fusion cells. From May 8 onward, it was Bockris and his experiments, at Texas A&M that, more than Utah, kept cold fusion alive. No one could dismiss or disbelieve it, not even Pons and Fleischmann, without first dealing

with Bockris and the spectacular phenomena generated by his cold fusion cells. John Bockris may have been the most perplexing figure in all of cold fusion. By 1989 his list of 600-plus publications was sixty-three pages long. He also had at least a dozen textbooks to his name. Several physicists, however, who picked up his Modern Electrochemistry to prepare them for cold fusion noted that it made electrochemistry seem like a science with no rhyme or reason.®° And electrochemists would point out paper after paper authored by Bockris that was dead wrong. Bockris thought of himself as a theorist, but then he floated his experimental results in print as though they were nothing more than idle theoretical speculation. His scientific philosophy seemed to confuse theory with experiment and was hard to reconcile with either. “I am not a person who believes that at the beginning of a new discovery, you should look too hard at it,’’ Bockris said. “I think you have to be very

encouraging and very positive.” When Bockris was later accused of publicly misrepresenting his results in cold fusion, he would say, ‘‘The good hard science, honest reporting, and so on and so on, in two, three years we'll worry about. But right at the moment, we’re right in the creation phase. The world is being born anew in this last year, and fantastic ideas are being discussed and being created, and right now it’s not good to ask too many questions.” Steve Feldberg may have put it best when he suggested poetically that “someplace along the way [Bockris] made the transition from being a

276 § GARY

TAUBES

good scientist to being someone who thought he had direct word from God, to being someone who thought he was God.” Feldberg called it “a logical but dangerous path.” Bockris was born in South Africa and raised in England. He went to London University and got his doctorate in chemistry at Imperial College in July 1945. When his adviser left just as he received his degree, Bockris was asked to fill the position. Chemists were in short supply nght after the war, and Imperial, said Bockris, “‘had a terrible difficulty in

getting anybody.” Bockris immediately set about building a team, a talent for which he would later become renowned. Within a year he had ten graduate students working under him. He later called the Imperial years a “heady, wonderful time.’’ Bockris said that when the young Martin Fleischmann wrote to him in 1947 looking to sign on as a graduate student, he had to reply that he was ‘‘full up.”’ Fleischmann went to work for another chemist at Imperial, whom Bockris called ‘‘a lugubrious individual.” Soon Fleischmann was working with Bockris, who was all of two years older: ‘‘At this stage there was no comparison between us. He was just a junior.” Bockris regarded Fleischmann as “‘a brilliant chap,” but he had a few problems with him (it is hard, though, to imagine the scientist with whom Bockris did not have at least a few problems). Bocknis recalled that Fleischmann had ‘“‘this dreadful habit of what I call ‘blah blah.’”’ As Bockris told it, Fleischmann would respond to his pointed scientific questions, “Well, you know, John, that’s a very interesting question, most, most, most debatable, most debatable. You’re going to have to

solve the Schrodinger [equation] for that one, John.’ ”’ Said Bockris, “After several of those I gave up. . . . It was hopeless.”’ In 1960 Bockris left Imperial because, he said, progress was in short supply in Britain; the country was rooted in the past. He crossed over to the new world and took a position at the University of Pennsylvania, where he takes credit for having built the largest physical chemistry group in the country, with nearly fifty chemists working under him. In 1972 Bockris left again, this time for Australia. America, he said,

had become a “terribly afflicted place. Vietnam. Student revolution. Everything seemed to be collapsing. Standards were falling in the universities. Deans were begging us to go easy on the students. Blacks were being brought in in large numbers. Everyone was welcome in the university. Didn’t have to pass an exam. Bring ’em in.”’ Bockris found a sanctuary at the University of Flinders in Adelaide, a new school planned for 7,000 students. But he came back to America in

BAD

SCIENCE

@ 277

1978 because, he said, he missed being where the action was. He landed

at ‘Texas A&M, which he’d heard had a strong interest in solar energy. He said he sat down for an hour with the number-two man in chemistry ‘slinging a few equations around,” and the university hired him. Bockris had earned his fame during those heady Imperial days. He introduced the study of electrokinetics, which prescribes that understanding reaction mechanisms on electrode surfaces comes about by characterizing how the reaction rate varies with the experimental variables. Some

of Bockris’s less avid fans—Nate

Lewis, for instance—re-

marked that this was the last good work he did, although that may have been doing the man an injustice. In the 1960s he had plenty of original ideas but, as Steve Feldberg put it, a lot of times he missed the boat. He also became famous in electrochemistry circles for some notable disputes with Martin Fleischmann, in which Fleischmann was invariably right. It is difficult to say when exactly Bockris began propagating flamboyant errors in lieu of good science. While he was still at Penn, for instance,

he claimed to have discovered an innovation that would herald the dawn of what Bocknis called the “hydrogen economy.” Newsweek once called it an idea of “breathtaking simplicity”: sunlight generates cheap, clean hydrogen fuel from water, which then powers everything from cars to utility plants. No more OPEC. No more pollution. And so on. The best estimates have always put the hydrogen economy at least fifty years away from commercial reality. And the process has yet to be proven to be thermodynamically or commercially advantageous, in part because the only good catalyst for reducing oxygen is platinum, which happens to be very expensive. In the late 1960s, Bockris said he had found just such a cheap catalyst in a tungsten-bronze alloy. However, Bockris’s researchers were synthe-

sizing his catalyst in platinum crucibles, and platinum will easily dissolve and recrystallize on other metals. This tungsten-bronze alloy turned out to be tungsten-bronze with a coating of platinum crystals. Rather than admit his error, Bockris claimed that tungsten-bronze is an excellent support for platinum and increases the catalysis action of the platinum, which is also not the case. This technique is known in some circles as the bait and switch. The tungsten-bronze alloy faded into history. In October 1982 Bocknis announced that he had stumbled upon a new inexpensive, and more efficient catalyst to split water into hydrogen and oxygen, which is to say he’d found another Holy Grail of the hydrogen economy. He reported that he’d used silicon electrodes which were

cheap enough, coated with a “‘secret catalyst.”” He claimed that his

technique was at least ten times more efficient than any existing method.

278 B GARY

TAUBES

This episode then ran like a prototype of cold fusion. Bocknis, for instance, made his announcement before any journal had accepted a paper on the subject. Bockris may not have written one yet. And he went public a few weeks after a UC Berkeley chemist made a similar, albeit considerably less dramatic announcement. An A&M spokesman then told reporters that the work “‘rivals such projects as the development of the atomic bomb in scientific importance.” Bockris himself said that hydrogen fuel would be selling for the equivalent of a dollar a gallon in the not-too-distant future. Asked how he had achieved the results he claimed, Bockris told the

Los Angeles Times, “I suppose that you'll have to trust me.” It may have been an injudicious comment to make to a reporter working so near Hollywood. The Times ran the article under the headline FUEL FROM WATER CLAIM TAKEN WITH A GRAIN OF SALT, an eminently appropriate angle for the story. Bockris also managed to get his picture in the papers. The photograph became celebrated in electrochemistry circles because in it Bockris is in the laboratory wearing his protective goggles upside-down. Perhaps it was his idea of a joke. Or maybe he just didn’t notice in the excitement. Bockris was busy at the time working on a $600,000 grant proposal to the National Science Foundation to launch his Hydrogen Research Institute.*? Bocknis turned out to be dead wrong about his silicon electrodes and secret catalyst. The system had been studied definitively by Al Bard, by Mark Wrighton at MIT, and by Adam Heller at Bell Labs, to name a few. Bocknis had overestimated its efficiency by ignoring much of the energy feeding the system. Seven years later this kind of error would sound vaguely familiar.*? In 1984 Bockris discovered yet another Holy Grail, this one in solar cell efficiency. “In the solar hydrogen economy, I am the man,” said Bockris. He claimed over 100 percent conversion of sunlight to electricity using a material the properties of which his colleagues believed were well understood. His claim violated at least one law of thermodynamics, probably two. Al Bard’s group at Texas reproduced the work but not the result, and Bard published a paper correcting Bockris’s misconception. The abstract said: ‘“As opposed to a previous report, quantum efficiency values above 1 were not obtained.” And so it went, paper after paper.®* As the cold fusion affair was winding down, Bockris told The Dallas Morning News, “I’m one of very

BAD

SCIENCE

§ 279

few scientists who don’t regard science as the principal way to knowledger? Al Bard believed that Bockris’s errors were a reflection as much of the inadequacy of the publication system as they were of Bockris. Bard couldn’t understand “how he slips that stuff into print.” He said that if anyone goes out of his way to correct the mistakes, as his group had on the issue of solar cell efficiency in 1984-85, it gives the appearance of a controversy when there is none. “He doesn’t get called on it,” Bard said. “You say, what the hell, I’m going to ignore it, but then it sticks in your craw.” Perhaps Bockris simply neglected the crucial experimental stage of looking for ways in which he may have been fooled and proceeded instead directly to publication. Or perhaps he believed he was unfoolable. Bockris was known to be so enamored of his theories that often his graduate students, in order to earn their Ph.D.’s came under pressure to procure experimental results that confirmed them. Ramesh Kainthla, Bockris’s senior researcher for five years, said he saw this problem develop a number of times; sometimes the graduate students gave in, and other times they left.® After March 23, if a cold fusion revolution was in the works, Texas

A&M University seemed to have the ideal combination of talents to lead the avant-garde. Bockris ran his own electrochemistry lab, and John Appleby ran the Center for Hydrogen Research and Electrochemical Studies (previously Bockris’s Center for Hydrogen Research back when Appleby was Bockris’s funding agent at the Electric Power Research Institute). Appleby was an expert in hydrogen fuel cells. When he signed on

at A&M

in 1986,

he hired Supramaniam

Srinivasan

and Oliver

Murphy, who were both former students of Bocknis, as Fleischmann had been. When

cold fusion broke, EPRI,

the research arm

of the utilities

industry, was already funding Appleby and Bockris, not to mention Chuck Martin. Among the three, it was spending some $150,000 each year on fuel cell research. EPRI then instructed these three to divert whatever funds were necessary for cold fusion. Rocky Goldstein, who was the A&M project manager at EPRI, said “there was an open pipeline.”” Tom Schneider, who was coordinating the cold fusion program at EPRI, told the administrators at A&M that it seemed “almost preordained” that, if cold fusion were real, A&M would lead the way. Quite simply, A&M had “‘the largest, most distinguished group working on the

280 B GARY

TAUBES

problem, anywhere in the world.”’ And, of course, Chuck Martin was

plugged directly to Pons, and Bockris to Fleischmann, giving A&M the edge on inside information, if any was forthcoming. Bockris had immediately made cold fusion the order of the day in his lab. He seemed to accept its validity without question. When he phoned Fleischmann

on March

25, his first words were

“My God, Martin,

you’ve done it this time.” Fleischmann replied, “Ahh, I hope it’s right, John.” Bockris’s response was ‘“Well, of course it’s right.”

Whatever the justification for this faith (“I reasoned that Martin Fleischmann wouldn’t make a fool mistake,’ Bockris would say later),

Bockris quickly directed six of his researchers to build cold fusion cells. They included Bockris’s senior scientist, Ramesh

Kainthla, an Indian

physicist via Australia; Marek Szcklarczyk, a postdoc from Warsaw; Jeff Wass, a Belgian graduate student whose support Bocknis cut off once he began writing his thesis; Guang Hai Lin, a theoretician from China;

Omo Velev, a Bulgarian with a master’s degree in physics; and, finally, Nigel Packham, a British graduate student who also served as Bocknris’s general factotum. Then Chuck Martin announced on April 10 that he had discovered excess heat. Packham’s wife, who was a local television news producer, told Packham, who in turn alerted Bocknis. Bockris said he was under the impression that the A&M electrochemists had an unofficial pact, a onefor-all-and-all-for-one arrangement, and would call the announcement,

even though Martin was wrong, a “‘dirty act’? and a “double cross.”’ Bockris appeared at the last moment to try to elbow his way on camera only to have Bruce Gammon, one of Martin’s collaborators, lock the doors on him. “I believe that was the first time I’d ever seen him,”

Gammon recalled. ‘He didn’t say, “Hello, how are you,’ or anything. He said, ‘Get out of my way.’ ”’ Or, as Martin put it, “John comes running up there to get in. .. . And John goes, ‘I’ve got to get by here, I’m about to go on TV.’ And Bruce goes, ‘I don’t know you.’ Pushes him out the door. Shuts the door.” A week or so later, John Appleby claimed that his group, which consisted of a single Korean postdoc by the name of Young Kim, had also generated anomalous excess heat. Kim was using a commercial instrument called a microcalorimeter. Appleby described it as a “‘a very sensitive instrument devised in the late ’60s or early ’70s to detect very small amounts of heat, originally for things like biological purposes.’’ Unfortunately, measuring small amounts of heat and measuring small variations in large amounts of heat are two entirely different problems. Addition-

BAD

SCIENCE

@ 281

ally, the cavity in the microcalorimeter was only a few centimeters across, so Kim had to run micro—cold fusion cells. Still, Appleby said the device was fine for cold fusion, and they found excess heat, so that was that.®

Meanwhile, Bockris’s group began building cells and doing calorimetry. Packham said they “really didn’t have any idea about nuclear chemistry,” and every time they tried to obtain a gamma ray spectrometer or neutron detector, Martin’s researchers beat them to it. Then again, their

calorimetry, as Packham said, was “‘primitive as hell.’’ They were having trouble competing with either Martin or Appleby, both of whom appeared to be doing sophisticated calorimetry. Both had announced they had excess heat. When they heard about the tritium announcement by Glen Schoessow and John Wethington, the two retired nuclear engineers at the University of Florida, Bockris and his researchers went looking for tritium. Said Packham, “‘It was like a race to see who could geta tritium counter first.’” The nuclear engineering lab at A&M had one, and Peter Lee, a Chinese graduate student, who knew how to run it. All Bockris’s

people had to do was walk over samples of the electrolyte from their cells. All science should be so simple. On Friday, April 21, Packham took samples from two or three of their sixteen cold fusion cells and brought them to Lee. After the weekend, Packham received the results of their very first tritium assays. Lee asked him what kind of source he’d put in. As Packham recalled, “I said, “There isn’t any source in there. It’s just electrolyte from the cell.’ And Lee said, ‘Well, there must be a source in

here. You have ten to the sixth counts per minute.’ At which point my jaw dropped. [I] walked back to the lab, asked Bockris for a meeting, had everybody together, and said I just picked up the tritium results. Instead of putting in the exponential, I just put it in longhand, and his eyes got wider and wider as I put this huge number down.” On April 27, Irwin Barton, a state representative, stopped by A&M to see about state funding for cold fusion. Appleby then announced publicly that he had excess heat. Quite a few members of the state government were apparently thinking, like the Electric Power Research Institute, that if cold fusion were real, Texas was the place to do it; and Bockris wanted A&M to join forces with Al Bard’s group in Austin to convince the state that it was a worthy cause. Bard and Ken Hall, who was both deputy director of the Texas Engineering Experiment Station and A&M’s associate dean of engineering, wrote a proposal for funding. But the legislature eventually decided against allocating any money.

282 8 GARY

TAUBES

This was, at least in part, because Bard had begun to doubt seriously the validity of the science. Or as Hall explained it, “The people that they chose to put the most complete confidence in were very negative.” None of the surrounding doubts waylaid Bockris, who now had two fusion cells that registered a trillion atoms of tritium in each milliliter of sample. Jaws were dropping. This number was some 10,000 times background. Tritium can only be created by a nuclear reaction. If huge amounts of tritium were appearing in Bockris’s cells, either somebody had put it there or a nuclear reaction was taking place ina test tube. Of course, to create one watt of power from fusion, there would have to be a trillion tritium atoms created each second. But maybe all the tntium was created in one second, or maybe they had created less than a watt of power. Still, tritium is tnitium. Lewis called Bockris’s group when he heard about the tritium from Martin, and, as he had done with Schoessow and Wethington, cautioned

that the tritium could be faked by chemiluminescence. But Packham later took a sample to Los Alamos, and the experts there confirmed it was tritium. On April 28, Bockris set out to prove that tritium had indeed been created in his cells. Beginning around noon his researchers took a single fusion cell, one labeled A7, and boosted the current. Packham and Jeff

Wass then removed two samples from the cell, the first at noon and the second two hours later. They assayed the cells together because the results from the two earlier cells were so unbelievable that there was already talk that perhaps they had been spiked. Wass explained that they sampled the cells together, “‘very simply because we didn’t want anybody accusing one or the other of cheating.” Kainthla, the senior postdoc, then took two more samples, one in the evening and the second near midnight. Bocknis hit the jackpot with A7. When the four samples were tested, they showed the tritium level dramatically increasing with time. They began with background levels at noon, slightly above background in the second sample, and then rocketed to 5.0 trillion tritium atoms per milliliter in the evening and 7.6 trillion near midnight.*’ Later, when

the talk of fraud became

more

prevalent and direct,

Packham said that their cold fusion cells ‘“were under guard for that time, twenty-four hours a day, seven days a week.”’ Of the protocol of A7, he said, “‘Four people were standing there the whole twelve hours in front of the cell when the samples were taken.” Packham argued that placing the tritium in the cell illicitly would have required ‘‘a gang of four.’’*

BAD

SCIENCE

§ 283

Kainthla, however, who took the two hot assays, said that none of the

researchers was guarding the cells. “If you think people were watching cells all the time,”’ he said, ‘‘that’s not true.’ Rather someone was in the lab, ‘“‘and once in a while he came and checked that the current was

passing through the cells and nothing unusual was happening.’’®? Kainthla later explained that the act of spiking would take all of five seconds. Remove the stopper from the top of the sophisticated test tube, stick in a syringe filled with tritiated water, and squirt. If the cells were in the calorimeter, it might take fifteen seconds, because a Teflon top

would have to be unscrewed. “If you want to do some mischief,”’ said Kainthla, “‘you don’t need a couple of hours. You can do it in a very, very short period of time.” Still, it was A7 that made Kainthla “‘more of a believer.” After all, he took the two last samples, and these had the

tritium. The following Monday, May 1, Packham found that three more cold fusion cells were newly endowed with huge amounts of tritium. That weekend, Bockris left for Los Angeles to attend the Electrochemical Society meeting, and yet another cell in a lab run by Kevin Wolf at the A&M Cyclotron Center, became hot.*? Wolfs background was equal parts nuclear chemistry and nuclear physics, which sounded perfect for cold fusion. Bockris had been delighted when he joined the research effort, because Wolf at least was not negative about cold fusion. This new hot cell had also been built by Bockris’s team. This cell, like the famous A7, was another lucky one for A&M. Packham and his colleagues took samples of the electrolyte from the cell on May 1, 5, 6, 7, 13, and June 6. Del Lawson, who worked for Chuck Martin, checked

the samples for tritium, although not until June. The sample from May 5 was negative, but the May 6th sample was hot, and the May 7 sample, slightly hotter. The samples from the 13th and June 6th showed no more tritium than on the 7th. The cell, however, had been surrounded by Wolf's neutron detectors when the tritium appeared and no neutrons whatsoever were detected coming from the cell. This was especially curious because when tritium is formed in deuterium fusion it happens with a healthy amount of kinetic energy—one million electron volts. The tritium, in other words, comes together with a sufficient kick to cause it, one time in every 100,000, to fuse with any deuterium atoms that might lie in its path. This tritium-deuterium fusion reaction supplies the “‘boost” in hydrogen bombs, which is why tritium is in such high demand in the nuclear weapons industry. The reaction had been studied extensively and has

284 0 GARY

TAUBES

nothing to do with cold fusion. The result of this tritium-deuterium fusion, among other things, is one 14 MeV neutron for each fusion, or, in the case of Bockris’s cell, an estimated 10,000 neutrons per second.

Wolfs detectors did not register 10,000 neutrons per second. They detected none at all. No one in Bockris’s group seemed to find this lack of secondary neutrons particularly suspicious. Nonetheless, the absence of neutrons strongly implied, if one had any faith in nuclear physics, that the tritium was not formed in the cell at all. Rather it must have come together, so to speak, elsewhere and entered the cell in a more circuitous

manner. In private, Bockris might casually mention the possibility of fraud— as he had said to Steve Feldberg, maybe someone spiked his cells. Stranger things had happened. In public, in Los Angeles, he professed to have no doubts about the tritium. It was iron-clad evidence for the existence of cold fusion. And the proponents of cold fusion would believe him. As Hugo Rossi would later say, Bockris’s tritium seemed to have clinched it.

65

SALTLAKE

ADMINISTRATORS.

CITY.

THE

PRESS

After the ordeal of Los Angeles, Stan Pons took a few days off and went with his wife to the Napa Valley. He called Rossi, who had been concerned about his state of mind, and said that everything was fine and that he’d gotten the chance to unwind. Back in Salt Lake City, the local press shrugged off the Los Angeles fiasco as though it had never happened. Martin Fleischmann’s confession that the gamma ray spectrum was worthless was not worrisome because they had, in effect, known that all along. Then, Pons had reported a burst of 5000 percent excess heat, which is the kind of huge number the press and administration could understand, meaningless as it might be. And why should they believe Caltech or MIT any more than John Bockris, who said that fourteen labs had confirmed, or Bob Huggins-Stanford, who told The Salt Lake Tribune that “‘it’s hard for [Nate Lewis] to see the effects because he is doing it wrong.” Then, too, Stanford and Texas A&M represented not only confirmation but competition. Both schools had applied, or would apply, for patents.”!

BAD

SCIENCE

@ 285

The Tribune led off with the headline u. TEAM UNVEILS NEW FEATS and said that Pons and Fleischmann ‘“‘had reported even levels of heat coming from their controversial experiment.”’ The News reported Los Angeles as a failure because it did not quiet the

FUSION higher Deseret critics.

Then it listed eight items of ‘‘new evidence presented in L.A.,” including, among other equally incomprehensible statements, that Pons and

Fleischmann had reported ‘“‘bursts of neutrons and other radioactive particles, usually ignored by physicists.” The Washington Post described the Los Angeles meeting as a disaster for cold fusion, claiming that Fleischmann and Pons were “facing a now almost overwhelming consensus among scientists that . . . cold fusion claims were false.” This prompted the Tribune to report the Post comment only in the context of how this kind of negativity would further frustrate any national response to the Japanese threat. Art Kingdom, an aide to Congressman Wayne Owens, told the Tribune that the political momentum had slowed not because of the bad science exposed in Los Angeles but because of the bad press that resulted. The Washington Post article, said Kingdom, was “‘being taken in Washington as further proof that things are not going Pons and Fleischmann’s way.” Two days after Los Angeles, Chase Peterson lectured on the “‘Politics of Fusion”’ at the university’s Hinckley Institute of Politics. He rationalized the disbelief away as the product of xenophobia and paranoia. “‘It’s less believed,” he said, ‘“‘because it didn’t happen in one of three or four

major research centers.”’ And, of course, it represented “‘a clear collision between small and large science.’ He concluded that they must not let such negativity deter them from their forceful march into the future. The News quoted him saying, “Have you noticed how quiet the Japanese are about this? . . . They may already be ahead of us.”’ Peterson and Jim Brophy then appeared before the state legislature arguing for the release of the $5 million, which still depended on confirmation. Peterson told the legislature that there had been “no disproving data”’ of Pons and Fleischmann’s experimental claims. He must have been convincing, because a local state senator told the Tribune that he was “still all warm and fuzzy” about cold fusion.”

286

B GARY

TAUBES 66

SALTLAKE

CITY.

PHYSICISTS

On May 9, Michael Salamon, a Utah physicist, went to work with Stan Pons in his basement laboratory. Although the majority of the local physicists were unified in their embarrassment over cold fusion, Salamon had a certain sympathy for Pons’s position. Back on March 31, when Pons was still open to help, the chemist had asked Salamon to make nuclear measurements in his laboratory. Now, the day after the Los Angeles conference, Salamon finally had his nuclear detection equipment ready to go, which was when the problems began. Salamon remarked later that he promised himself he would be completely objective: “‘I have to be. I can’t go in there with any bias.”” He added that he was skeptical, but wanted nothing more than to find evidence of fusion. “‘It’d be one of the most exciting things I’d ever do in my career,” he said, “‘But you’ve got to remember what you have,” which is to say nuclear fusion in a test-tube. The thirty-eight-year-old physicist claimed ironically that he could remain objective because of his history as a physicist who had never done an experiment with a positive result. He was not averse to working ona billion-to-one long shot if he thought it would illuminate matters. Salamon’s experience with negative results began with his graduate studies at UC Berkeley, where his adviser was Buford Price, who had just

come to believe that he had discovered an elementary particle known as a magnetic monopole. Price claimed the discovery on the basis of one peculiar observation made by a detector mounted in Skylab. He was eventually convinced by Luis Alvarez among others of his error, and Salamon witnessed first hand the danger of going public and being wrong. The lesson was well taken. Salamon then obtained his doctorate on the basis of an experiment designed to detect atoms of anti-iron, which are iron atoms made out of antimatter instead of matter. It was an interesting experiment; however, it came up empty. He also spent six months laying out a huge monopole detector in an empty lot at the Lawrence Livermore Laboratory. It was intended to sit undisturbed for decades and may or may not have come up empty. Several years later, the Lab put up a building on the lot and, not knowing what the detector was, the workers tossed it in the city dump. Then there was a memorable expedition in August 1987 to Prince Albert, Saskatchewan, this time looking for anti-protons in cosmic rays.

BAD

SCIENCE

#8 287

After a year of preparation, Salamon and his colleagues lofted a detector in a high altitude balloon to 130,000 feet. Salamon returned with pneumonia and another negative result—no cosmic ray anti-protons. More recently, he had an ongoing experiment with his Utah collaborators called the Fly’s Eye, a cosmic ray observatory built out in the desert. The Fly’s Eye had recently received publicity as an example of good science done at the University of Utah and may or may not have observed some new and bizarre phenomenon. Salamon doubted it: “I think those are negative results, too.” In the spring of 1989, Salamon had three projects in the works to keep him busy enough without cold fusion. But he was not going to walk away from it: “I am the only one in the world who has the opportunity to make the measurements in Stan’s lab. Besides it’s a great excuse to learn live nuclear physics.” Salamon’s partner in the endeavor was Ed Wrenn, a nuclear engineer

who ran a radio-biology lab at the U. Salamon and Wrenn mounted a sodium iodide gamma ray detector beneath the table that held Pons’s fusion cells. This would observe gamma rays twenty-four hours a day. Salamon also built a neutron detector of uranium 235 foils sandwiched between plastic nuclear activation foils. This was the kind of detector Buford Price had used for his monopole. Salamon described it as a “‘very nice, sensitive’ neutron detector that relied on fission, rather than fusion

to work. Most neutrons trying to pass through the uranium foils would hit a uranium 235 nucleus, causing it to fracture into two smaller nuclei. These fission fragments would then leave pits in the plastic. ‘““You look in a microscope,” said Salamon, “‘and you look at the little pits in the plastic.” Count the number of pits and that gives the number of neu-

trons. The device was elegantly simple. Its only drawback was that uranium 235 was a controlled substance, strictly safeguarded by the Nuclear Regulatory Commission. Anytime Salamon wanted to use this detector, he had to oblige NRC regulations. This meant, for instance, that he couldn’t just stash the foils under Pons’s cells for weeks on end, but had to leave them in the nuclear engineering department and then hustle them over if Pons, Marvin Hawkins, or Mark Anderson told him they

had a cell actually generating excess heat. By the time Salamon’s equipment was ready to go, Pons told him that he had shut down all his active cells and it would take a couple of weeks to start up new ones. Only a week later, however, Jim Brophy told The Deseret News that Pons had seen bursts of one hundred times excess heat,

lasting from a few hours to days. Brophy seemed to be implying that Pons

288 BE GARY

TAUBES

had witnessed these episodes recently. Salamon later said he asked Pons about this episode, and Pons said the newspaper account was inaccurate. It was all very mysterious. Meanwhile, journalists had already begun asking Salamon about his data, and he was beginning to wonder what he had gotten himself into. Salamon, who has the spirit of a New York intellectual, said it reminded him of a poem by Catullus, which began: Lost your mind, Ravidus, you poor ass,

landing smack into one of my poems like this? Is some god getting you into trouble Because you didn’t say your prayers right? Or are you just out to get talked about? What do you want? To be famous, never mind how?

;

67

THE

PRESS

On May 18 the results of Richard Petrasso’s gamma ray analysis appeared in Nature to little notice, even in Utah. This was not surprising, for the

journal had run the article under the thoroughly undramatic headline “Problems with the [gammal]-ray spectrum in the Fleischmann et al. experiments.’’ Petrasso demonstrated conclusively that Pons and Fleischmann’s primary evidence for fusion, the gamma ray peak, was suspiciously bogus, but he presented his case in so unemotional and scientific a manner that it was easily ignored. The Deseret News ran a couple of paragraphs on the article and summed up, ““The MIT scientists offer no explanation for the energy measurement except that it may be ‘an instrumental artifact’ unrelated to gamma ray interaction.”’ In comparison, The Salt Lake Tribune’s lead story that day quoted Chase Peterson saying that if they didn’t get cracking on cold fusion in the next two months, they’d be run out of business by Stanford and the

rest of the competition. ‘““Those who have successfully repeated the experiments,”’ said Peterson, “are now saying, “They may have stumbled upon something out there [in Utah]. But they don’t have the weight to carry it. The big boys can doa better job.’ ”

BAD 68

SALT

LAKE

CITY.

SCIENCE

§ 289

ADMINISTRATORS

Carlton Detar, who was one of the more vehement cold fusion detrac-

tors among the Utah physicists, had been engaged since March 23 in a sporadic assault on the Utah administration. Hoping to bring the cold fusion bandwagon under control, if not to a halt, he had been sending Jim Brophy memos expressing his opinions about cold fusion, but Brophy never responded. When Hugo Rossi was appointed head of the Utah cold fusion effort in early May, Detar began his assault on Rossi. The appearance of the Petrasso paper prompted the first foray. “The questions raised in this paper,” Detar wrote to Rossi, “are extremely serious and represent a charge of incompetence to say the least.’ Detar went on to enumerate the various reasons why the university’s reputation might be forever damned and ended with yet another variation on Pascal’s wager: ‘Given what has been established scientifically about cold fusion so far, we might as well be reallocating resources and reassigning manpower to find the pot of gold at the end of the rainbow.”’ Rossi responded that he had similar concerns and had concluded that Detar’s letter deserved “‘a serious response.’’ He wrote: The fact is that I believe that Fleischmann and Pons have made a remarkable discovery which—for whatever reasons—they felt compelled to disclose prematurely. At first I was skeptical of the science and the hype. . . . But, as I listened to the explanations of their work, and its history ... as I talked with others who had confirmed this energy production both in heat and radiation, I began to accept that not only was it possible, but that it was real. As I witnessed the virulent reaction of the leaders of our national scientific enterprise to a challenge from a discovery which occurred out of the system, I began to understand the non-scientific aspects at a deeper level, and began to accept the necessity to become a part of it. Isn’t it extraordinary to argue, on the basis of failed experiments that claimed successes had to have been failures? Isn’t it strange that Pons and Fleischmann are viciously attacked as incompetents while the mounting evidence of reproducibility is not acknowledged? What is going on scientifically? What is going on politically? The cost to the University in resources, fiscal as well as human, is much

less than it appears. And the loss in stature and reputation, no matter how adverse the outcome, will be small and temporary if we maintain our conception of who we are. I am convinced that the science will hold up.

290 B GARY

TAUBES

Detar countered with a second memo, in which he included a copy of a preprint written by three German chemists, G. Kreysa, G. Marx, and W. Pleith, who had carried out a critical analysis of the PonsFleischmann calorimetry similar to what had been done at Caltech. Both research teams, as Detar pointed out, claimed that Pons and Fleischmann’s heat measurements could be explained without invoking nuclear fusion. Rather, they could all be accounted for by chemical reactions, sloppy calorimetry, impossible hypothetical situations, or all three. Rossi sent a copy of Detar’s memo to Pons, who responded with vitriol two weeks

later; he called Detar’s criticisms

“ludicrous”?

and

“really depressing,’ adding that Detar’s ‘ ‘‘serious objections’ are not worth addressing further. They have no basis in fact!”’ Rossi then wrote Detar pointing out the “unscientific bias’ in his thinking. He accused him of accepting Nate Lewis on faith, even though admitting he had not yet seen his paper, while rejecting ‘your colleagues’ without even walking over to their lab to talk with them!”’ He concluded:

I have made no claims of primary knowledge of any of this theory or experiment, and it would be ludicrous for me to do so . . . your insistence that I be ‘scientifically’ responsive displays an arrogance that I find deeply troubling. Do you want me to tell you about my experiments? Have you told me about yours? Or do you want me to confess that I have no scientific basis—and if so, will you do the same?

Rossi’s faith always seemed at odds with his obvious intelligence (though as a mathematician, he had a limited grasp of physics and experimental techniques). Rossi never stopped trying to do the right thing, yet his actions, out of context, made many scientists question his sanity. In context, he was a victim of fusion fever. Rossi, who was in his mid-fifties, had grown up in New York, the son

of an Italian anarchist. His father had been deported back to Italy as a subversive, then reentered the United States illegally after what Rossi called “the assassination of Sacco and Vanzetti” and spent the rest of his life as an illegal alien, writing for an Italian anarchist journal until the 1970s. Rossi said that from his father he received a strong work ethic “and a real interest in trying to think things through and analyze things.”’ Rossi did his undergraduate work at City College in New York, which seems like the appropriate place for the son of an immigrant anarchist; then he went off to MIT, where his thesis advisor was Isadore

Singer, a world-renowned mathematician. Singer later said that Rossi

BAD

SCIENCE

# 291

was probably the best student he had ever had. From MIT, Rossi went to Princeton, where he did excellent work in mathematics, and then to

Brandeis, where he stayed for eleven years and became chair of the department. In 1974 he moved to Utah because he was, he said, “‘sick

of academic pretense” and looking for a better life for his family. Between 1974 and the dawn of cold fusion, Rossi and his friend Joe Taylor helped build the Utah mathematics department into a first-rate program. He became dean of the College of Science in 1987, when Taylor stepped up from dean to academic vice president. Singer, who had remained

close to Rossi, said that Rossi did not have a natural

inclination for administrative work but was so solid that colleagues would turn to him for help. Rossi would then “‘put his shoulder to the wheel,’ which is what he did in cold fusion.

By May the Utah administration hoped to launch a major cold fusion project capable of sorting out all the unknowns as quickly as possible. If the federal government wasn’t going to help, then Utah would do it alone. It was clear now that Pons had no inclination to collaborate and little inclination to share information. Rossi was the only member of the administration, other than Jim Brophy, who had any kind of relationship with Pons. Pons seemed to trust Rossi, and Rossi had come to like Pons.

He found his personality intriguing and was taken by his audacious sense of humor. Perhaps most important, Pons had convinced Rossi of the validity of cold fusion.” As Taylor would say, Pons exhibited an almost uncanny ability to turn skeptics of cold fusion into believers. Simply put, the closer you got to Pons the more you were willing to believe. “He might look like a jerk at a distance,” said Taylor, “but up close he was almost hypnotic.” Rossi became the natural choice to head the program. According to Taylor, ‘Chase [Peterson] never should’ve been involved in the first place. And Brophy was speaking to the press all his line of bullshit. The faculty was incensed about how this was being portrayed as a sure thing [and] a solution to the world’s energy problems. We had to get [Brophy] out of there. The idea was within a few months, if this didn’t collapse,

we would hire a permanent director. We just needed someone there, and Rossi, [in] his position of dean of science, was doing a lot already.” Rossi was in a no-win situation: he would have to please Pons and Fleischmann on the one hand, his faculty on the other, and he needed

a third hand to deal with the community of scientists at large, who'd pegged him as a bad guy as soon as he appeared in the press defending

the cold fusion work. Rossi took the job, however, and kept at it because

he believed that it had to be done. He wouldn’t allow himself to be

292 I GARY

TAUBES

swayed by the difference in styles between, say, Nate Lewis and Bob Huggins. To Rossi, Lewis was grandstanding, ‘‘acting inappropriately,” and Huggins was simply trying to get his point across. Rossi was still reading the national and scientific press, and he was in contact

with friends at Princeton

and MIT,

which

should have

il-

luminated the issue for him but did not. “I was taking one hell ofa lot of abuse,’’ Rossi recalled. “I am somebody who left the whole Cam-

bridge complex to come out to Utah not because it’s God’s territory but because I rebelled against the smug sureness. And that’s what I kept seeing in all this. Maybe it was just wishful thinking. God, I want those smug assholes to be wrong. Those people who were just so damned sure of themselves.’’ Rossi believed Peterson felt much the same way. Rossi knew about John Bockris and the Texas A&M tritium within a week of its discovery. As Bockris kept calling Utah with news of his latest tritium eruption, Rossi took to walking down to Pons’s lab and asking Mark Anderson what Pons had in the way of tritium. Anderson told him that they had seen three, maybe five times background. This helped to comfort Rossi, who may not have understood that this tritium accumulation would happen naturally during electrolysis. Rossi saw Huggins as their “‘savior,”’ but he didn’t learn about the flaws in Huggins’s research until June. When he confronted Huggins with the criticism, Huggins shrugged it off in his inimitable way; he said it was irrelevant, inconsequential. Rossi believed him. By the time Rossi regained his objectivity, some six months later, he had so alienated many of his faculty that they hoped to prevent him from returning to his position as dean. The Utah physicists said that the embarrassment of having the former cold fusion czar as their dean would be too great. Rossi eventually won back his faculty’s support, but it took a year to do it.

69

SANTA

FE,

NEW

MEXICO

Ryszard Gajewski first floated the idea for a Department of Energy— sponsored cold fusion conference on April 11, the day before the Dallas American Chemical Society meeting. He wanted to hold it in either Los Alamos or Brookhaven, but settled on Santa Fe because of the difficulty

BAD

SCIENCE

# 293

in allowing “‘foreign nationals” to visit Los Alamos, a high-security lab. After Baltimore, Rulon Linford suggested to Gajewski that the meeting had become irrelevant: ‘‘everything that could have been said, already has been.’’®> It was Gajewski’s show, however, and he believed in

what Steve Jones, and maybe even in what Stan Pons and Martin Fleischmann, had reported. Gajewski certainly felt an overwhelming responsibility to cold fusion. After some discussion, he and Linford decided that the field could still benefit from a meeting in which “‘physicists and chemists could get together without acrimony.” The Workshop on Cold Fusion Phenomena began on May 23. This was the sixth major meeting on the subject, not counting the congressional hearings, in two months, which works out to one meeting every week and ahalf. Fifteen hundred scientists were expected to attend, but the workshop drew barely 500, which indicates that the social phenomenon of cold fusion had peaked and had made the downward turn. Jim Brophy told The Salt Lake Tribune that Pons and Fleischmann, who had been invited, weren’t attending the meeting because they were “better off spending the time doing the research so they would have something to report.”’ This sounded uncharacteristically sensible. (Pons added that his research was ‘‘going fantastic’? and that his critics are “‘soing to have to eat a lot of crow.’’) The meeting convened at the Sweeney Center in Santa Fe. The press was relegated to the gallery, looking down on the proceedings, which was not unlike the Romans watching the lions and the Christians have at it. The reporters were not allowed to ask questions from the gallery, but the workshop organizers had arranged press conferences following the sessions. This provided the various cold fusion researchers with a forum to talk to the press without their peers around to contradict them. Those researchers with positive results, however dubious, found it

surprisingly easy to get press coverage, as their results constituted the only real ‘news’ at the conference. MIT, Caltech, Bell Labs, Brookhaven, Yale, Los Alamos, Princeton, the University of Michigan, Sandia National Laboratories, Argonne, the Max Planck Institute in Germany, the

University of Tokyo, and the Chalk River Nuclear Laboratory in Canada were among the institutions reporting negative results and/or explaining the positive results of Pons, Fleischmann, Jones, and others in the context of bad calorimetry and bad nuclear physics. But they were all old news. Even Science, which constitutes a technical journal in this field, managed to report on the workshop without once mentioning those groups that had overwhelming negative results. What was news was what was new, and that was what supported cold fusion.

294 @ GARY

TAUBES

Without Pons and Fleischmann present, Texas A&M was left to carry their torch. John Appleby started it off, addressing the participants with his detached air of jowly, British academia.

‘““We do not yet have any

evidence that would stand up in court,” he said. Then he reported his excess heat results and added that Rockwell International had analyzed two of their heat-generating electrodes for helium and found none. The helium was the one exception to the new-news-is-good-news rule. The helium had been on the reporters’ minds since Los Angeles, when Pons and Fleischmann had been so cagey about it. Now Appleby had claimed excess heat and found no helium. Appleby said he was disappointed but added that it was no surprise because they hadn’t detected any neutrons either. Appleby also reported that they hadn’t found any tritium of note in their cells. He did report on some of John Bockris’s tritium results. This was peculiar because Appleby later observed that he had bumped into Bocknss that very night and said to Bockris, “Look, concerning this tritium, are you not sure that somebody hasn’t been spiking your cells?” Bockris replied that he’d taken all the necessary precautions, which wasn’t true, but it apparently mollified Appleby. Curiously enough, it was not Bockris who presented the bulk of the tritium work, but Kevin Wolf, the nuclear chemist, as Bockris put it,

ninety-nine rungs beneath him in the hierarchy. Wolf, who had a good reputation with physicists, said he presented the data ‘‘so that it might be more scientific in nature.” It seems several members of the Bockris group, including Bockris himself, had suggested that Wolf front for the work. Wolf later explained, quite simply, that ‘““Bockris had trouble keeping his facts straight.” Wolf seemed to be the antithesis of Bockris, if for no other reason than

that he was as American as could be. Although he was in his forties, he looked ten years younger. He was tall, slender, boyishly handsome, and

spoke with a slight drawl. He had even been arace car driver in his youth and given it up, he said, when he suffered through a bout of common sense. ‘Do you want to be steak Diane? was basically the question,” he said. Wolfs doctoral thesis was on nuclear fission, and he did heavy ion experiments at Berkeley for a while with Steve Koonin. The two had a mutual, if lopsided, admiration society: Wolf said Koonin was the best nuclear theorist in the world. Koonin said Wolf was one of the better people in his field, although not in the top five. Wolf then spent a decade at Argonne and moved down to A&M seven

BAD

SCIENCE

8 295

years before cold fusion. Lately he had been working on a heavy ion experiment at Brookhaven, although once he set in on cold fusion, he spent less and less time on his physics responsibilities. When cold fusion broke, Wolf opened his doors to anyone who wanted a cell counted for neutrons. Del Lawson, who worked for Chuck Martin, brought cells in,

then Nigel Packham came in with some of Bockris’s. Wolf ran Bockris’s cells for three weeks and was ready to quit when he detected neutrons from two cells, and Packham found tritium in one of the cells also.

Wolf had been hesitant about reporting the tritium in Santa Fe. He had been on what he called his ‘contamination kick.”’ All of the hot cells had used palladium from the same batch of one-millimeter-diameter wire, which implied that the tritium may have been in the wire all along. After he left for Santa Fe, one of Bockris’s researchers called to report that a new cell had come up hot. It had a three-millimeter palladium wire, which meant a different batch of palladium. So much for contamination. ‘You might interpret this as being somewhat suspicious,” Wolf said, but he reported the observations of tritium nonetheless: seven cold fusion cells with as much as 100 trillion tritium atoms—stunning levels. After presenting Bockris’s results Wolf broke with him for good, when Bockris chaired an evening session and tried to repeat the Electrochemical Society faux pas of excluding any speakers with negative results. After some vehement and incredulous arguments from the scientists present, Bockris agreed to accept a short talk from Dick Garwin. Mike Nauenberg, who was there, said he’d never heard a chairman like him:

“He only wanted to hear positive things. Totally obnoxious. I said to myself, Who the hell is this character?” Wolf told reporters the next day that Bockris was

‘‘an embarrassment,’

and that was

the end of the

collaboration. Nonetheless, A&M, via Appleby, had found heat and neutrons but no

helium, and Wolf was reporting that he had observed a few neutrons and Bockris had found plenty of tritium. What was one to make of it? Who knew? Certainly not anyone who read the papers the next day. The Wall Street Journal headline over a story by Jerry Bishop read: TEXAS GROUP REPORTS MORE SIGNS OF ‘‘COLD FUSION”’ AT U.S. MEETING. Bishop concentrated on tritium: SANTA FE, N.M.—Researchers at Texas A&M University continue to find indications that “cold fusion’’ is taking place in palladium rods, supporting the controversial claims of two University of Utah chemists. The Texas scientists told a cold fusion conference here that as recently as

296

B GARY

TAUBES

this weekend, their experiments were continuing to produce excess heat or large amounts of tritium, a form of hydrogen that can be produced by the fusing of hydrogen atoms.

The New

York Times headline read CONFERENCE

ON FUSION TOLD

OF

FAILURE. The Times’s reporter, William Broad, who had’a:much more

skeptical turn of mind than his Journal colleague, concentrated on the lack of helium: SANTA FE, N.M., May 23—Scientists at the first Federal conference on low-

temperature fusion reported today that they had failed to find some byproducts that had been expected from the nuclear reaction. John Appleby, director of the center for electrochemistry at Texas A&M University, told the conference that a hunt for the expected by-products, two types of helium, in the electrodes of the cold-fusion cell had been unsuccessful.

Next, Steve Jones, whose routine was getting more polished with each conference, gave his report. Perhaps he now felt that the field of battle was finally his; after all, Pons and Fleischmann had not shown up, and

his neutrons were being taken much more seriously than the excess heat. Jones reported his neutron data and claimed that they were confirmed by Wolf, as well as a Los Alamos researcher, who claimed to have seen bursts

of neutrons, and several Italian groups. (These included the Frascati results, which would soon be recanted.) Jones repeated that his cold fusion promised no energy, so it had nothing in common with Pons and Fleischmann’s. ““You put one watt into our cell,” he said, “‘you lose a watt. You get a few neutrons. You get no heat.” Later Jones was challenged by Moshe Gai, a Yale physicist who had done variations on both the BYU and the Utah experiments with considerably better neutron detectors and had seen no neutrons at all. Gai said he had a background rate of one neutron an hour and didn’t see neutrons. He added that a French group working in a tunnel underneath Mont Blanc recorded a background rate of two neutrons every five days, and they didn’t see a neutron signal either. Jones responded that Gai hadn’t used the exact BYU prescription, which made Jones sound like Pons and Fleischmann. He said Gai used palladium rods. ‘““We explicitly say no rods,” he said. Gai denied that they’d used rods. Well, are [the electrodes] fused? Jones countered. Gai said he didn’t know. “I’m not sure you’re doing the right experiment,” said Jones.

BAD

SCIENCE

B 297

So Gai suggested that the only way he would believe Jones was to have him bring his cells to Yale and test them on his system. Later Jones said to the reporters, “I’ve agreed to help [Gai] set up experiments. We'll see neutrons. I’m confident if we do the experiment there, we’ll see the effect.” It made for good press. Jones seemed to be playing the audience. “Let’s challenge Pons and Fleischmann to go to Yale and do the expenment,” he said. Science, at least, bought it, noting that, ‘‘although there were continuing debates at the workshop about the reliability of the heat measurements,

most

of the

attendees

were

convinced

that

the

second

phenomenon is real—some tiny amount of fusion is going on.” The Wall Street Journal bought it too: ““The detection of low levels of neutrons reported here did, however, convince many researchers that fusion of

deuterium atoms is likely to be taking place in the palladium.” Jerry Bishop went on to report that ‘“‘most scientists agree heat is produced but debate continues over its cause.”’ In other words, Bockris and Huggins professed that the heat was too great to be chemical, and virtually everyone else professed it wasn’t, or simply said they were “puzzled.” Both Huggins and Bockris, according to the Joumal, “steadfastly refused to speculate on what is producing the excess heat,” which made the two seem downright cautious. In fact, Bockris would refer to the “alleged fusion,” then say he much prefered to call it the “Pons-Fleischmann effect.” Huggins-Stanford professed absolute faith in the results of his latest series of experiments. “Indeed,” he said, “‘there are very definite effects

occurring having to do with thermal effects of the system.” This, not surprisingly, is meaningless if one stops to think about it. Huggins claimed that his researchers used the right palladium, prepared in the right manner, which explained why he hada heat effect and no one else did, or no one else but Pons and Fleischmann and Appleby, who used different palladium, prepared in a different manner. It was also a nifty argument: nobody else does it right. Later some Los Alamos scientists asked Huggins if they could share his palladium, since he claimed 100 percent of his electrodes worked. Huggins said he was hesitant because he didn’t have much to spare, so they asked him if they could give him palladium, which he could prepare in his magical way and send back to Los Alamos for them to run. Huggins said he would like to, but it would be expensive and he needed outside support. Nate Lewis asked Huggins again if his Caltech group could borrow one of the Stanford cells so they could run it near working neutron detectors or autopsy the electrodes, which might reveal tritium

298 BP GARY

TAUBES

or helium that would have accumulated during the fusion process. Huggins said they were too busy. Maybe later. Huggins seemed averse to jeopardizing the only two electrodes in the country that produced consistent excess heat. He may have realized that if anyone could prove his rods did not emit neutrons or autopsied them and found neither tritium nor helium, his arguments would be compromised. Steve Koonin was one of many who said it sounded like a swindle.”¢ In three full days of talk, only Daniele Gozzi of the University of Rome reported seeing neutrons and heat simultaneously. Gozzi had apparently been worried about the possibility of a nuclear meltdown. He had set his instruments to shut down his cell should its temperature reach 80 degrees centigrade. Before termination, however, the cell temperature soared to 150 degrees. Simultaneously Gozzi’s detectors observed thirty-six neutrons. Of course, for the neutrons to have been commensurate with the heat, there should have been a billion times more, so maybe

it wasn’t fusion. Gozzi said he’d had only one such event, and no one was there when it happened. When he returned to his lab, the cell had bubbled over. It sounded oddly like Pons and Fleischmann’s meltdown. Science called the workshop the end of act one and said that it could be several more months before anyone knew for sure what was happening. Norman Hackerman, a distinguished chemist from Rice University and cohost of the workshop, summed it up by saying, ‘“We have reached no consensus. Those who didn’t believe in the phenomenon have not changed their minds. Those who do believe there is something there haven’t changed their minds.” A. J. Liebling might have said that this was not a reflection on the expertise of experts, but on the futility of flapdoodle.

7TOLA

JOLLA,

CALIFORNIA

While the Santa Fe workshop proceeded, Chuck Martin of Texas A&M was having an epiphany on a beach in La Jolla, California. Martin had learned back on April 12 that the excess heat he had announced in his press conference—an announcement that had been reported around the world—was simply wrong. However, he still had one cell, one experi-

BAD

SCIENCE

@ 299

ment, that may have had legitimate excess heat. And within a week of his announcement, the issue was confused by John Appleby at A&M, who was suddenly claiming that he had excess heat as well. Martin and his researchers had built the cells for Appleby and given instructions on how to run them. Martin said he “didn’t know what to believe.” So he went back to the laboratory and continued to work at reproducing the experiment. He thought it would be easy. Martin’s story may be proof that it is possible to make it back from the abyss if one works at it long enough. “So we set out to try to reproduce this one positive result,”’ he said, “which is admittedly compromised. And we started doing it again and again and again and again and again and again, and it didn’t do it. So we just kept getting more and more single-minded. ‘Goddamn it, why won’t this work, why won’t this work?’ We kept trying different things, doing the experiment slightly different ways. ‘At first we thought the key was turning the cell off and turning it back on. Pons is saying things like, You need to perturb the cell somehow to make it go. We thought that turning the cell off would be it. And we did that over and over and over again, and it just didn’t do it. Then we thought maybe there was something in the way we pretreat the electrode. We tried a whole bunch of different ways: thermal annealing, not thermal annealing, plain crunching, then we tried casting electrodes. We ordered electrodes from every conceivable source, and nothing else

came up. “T talked to Stan a number of times during this period. In fact, there was a time when I talked to him every day, sometimes twice a day. The thing is he couldn’t help me. And the checklist would change. Even early on I couldn’t get a recipe from Stan on how to make this work. He was very evasive. I thought lawyers, maybe. But it was a worrisome thing to me that I couldn’t get the information from him as to how to make this experiment work. I talked to Martin [Fleischmann]. I talked to Brophy. And yet, they would say, ‘Just electrolyze it, you'll see it, you'll see it.’ And

all that time, we

didn’t have a computer,

so we

sat there and

tweaked knobs. It was horrible. It was absolutely horrible. “’d come home on a Friday night, and I’d go to Deb [my wife] and say, ‘God, it’s just not working, what am I going to do?’ And every once in a while I’d think, ‘God, I hope everybody doesn’t think I’m an idiot. I hope everybody doesn’t think I jumped on this bandwagon, and I’m going to get stained with this brush.’ Sometimes that would become an irrational phobia. .. . “T’'ll tell you the big thing that happened here was, my father-in-law

300 8 GARY

TAUBES

passed away right before the [Santa Fe] meeting at the end of May. The week after that, Deb had a meeting in San Diego. So I go out to the funeral, which obviously forced me to quit thinking about this problem. Then Deb said to me, ‘Chuck, go with me to this meeting. Step back from it for a week, just a couple of days. You’ve been working twentyfour hours a day. You’re a zombie. You can’t do good science this way.’ Initially I resisted her, and then I said, “You’re nght.’

‘So I went down with her to San Diego, and she went to the meetings every day, and I did nothing. It was during that period of lying on the beach, I said to myself, ‘Stan is wrong. Stan Pons is wrong. This stuff is not right. We had had some positive effects, but the vast majority were negative, and when we did very very careful experiments, they were all negative. Maybe not being able to reproduce it is the right result. For whatever reason, this thing might be an artifact. And I can’t go on living the way I’m living, and as much as I can’t believe that my friend Stan Pons is wrong and that my friend Stan Pons lied to me, those things are truese: Martin later observed that Pons and Fleischmann might have experienced something similar to what he went through, but on a much more grandiose and horrible scale. ““Thatis,”’ he explained, ‘“‘we saw an effect; we

tried like crazy to reproduce it and we couldn’t. And then we started to wonder if anything was real.”’ This reminded Martin ofa line in Foucault’s Pendulum, anovel by Umberto Eco: “‘I believe that you can reach the point where there is no longer any difference between developing the habit of pretending to believe and developing the habit of believing.’’ Martin suggested that Pons and Fleischmann had reached that point and crossed it about the time that he went public, perhaps in response to his announcement and several others that quickly followed. Martin noted that sometime in May he had called Pons to tell him that he was withdrawing his excess heat paper, which he had submitted back on April 9. Martin said, “Stan, you know that paper I sent to the Journal of Electroanalytical Chemistry?”’ “Yeah, I’m reviewing it,” Pons replied. “Stan,” Martin said, “I’m withdrawing that paper.”’ ‘““What? You’re withdrawing it?” said Pons. ‘‘Those results look fine tomes Pons knew the results were compromised, Martin explained, “but he

still tried to talk me out of withdrawing the paper.” Pons apparently saw Martin’s epiphany as a defection. Around this time Pons said to Hugo Rossi, ““Chuck’s gone wrong; they’ve gotten to him. He talks too much to Lewis.”

Book Three

THE TAIL OF THE DISTRIBUTION If science is to progress, what we need is the ability to experiment, honesty in reporting results—the results must be reported without somebody saying what they would like the results to have been—and finally—an important thing—the intelligence to interpret the results. RICHARD FEYNMAN, The Character of Physical Law

In the popular press we are always reading that “most scientists believe”? such and such. Who cares what most scientists believe? We

want

to know

what the BEST

ones

believe,

especially those in the best position to evaluate the topic at issue.

MICHAEL T. GHISELIN,

Intellectual Compromise: The Bottom Line



7

2

ive

=

4 y

AOD

\ ly ores

7



4

i

6%

aa

.

ict

*

yan

1.

vy

ro

£

an

i

a

vw ried

arene

Pion. f

wenn Pee

:

RG

yh ot sito ye

Ft

aoc

dine

eo

aw hoger a

re

se

TH cr ltick

if ain)



i ’>



;

Great

pity

te

mar

®

wes

re

ie FS

og wtA bees At tomtage ee ee cx

tard

atin ye Bly

ia?

»)

ote

ct)?

a

«iha

teh :

tox

e Rhee

Gia

eta

ee

evga oe age hele eee ree, Sect anak Pea Seerege

Font,

Wak



i’

het paMitie tm, «

cas

st



ieHOLT DBT SF RAabrace ea ,

oe Tne siheplioe

wmipwipefes®

ane Oat ’fiype on

ie

(eeeet, Cr

;

ree

Oe ae

ny Pat das rarored ie, ;

ee

ee)

ee

ee

Oy 4

eee

ae

a

Garay ave pre a

aS

te

oe._

nl can! ae

eee rf

ae

T ROCHESTER, SANTA

FE,

NEW

NEW

MEXICO.

YORK. SALT

LAKE

CITY

“If you’re a scientist,”’ John Huizenga has said, “‘you’ve got to be skeptical. Don’t believe anything straightaway, until somebody has given you some hard proof.’ Regarding cold fusion and related long shots, he believes that ‘“‘when something violates all that we know about nature, you better have some darn good proof before you start mouthing off about it.”’ Huizenga was a veteran of the Manhattan Project. His doctorate was in physical chemistry, but he had made a career out of nuclear physics, and when cold fusion broke he was a professor of both chemistry and physics at the University of Rochester in New York. He had been at the Dallas meeting on April 12 and had returned from it demoralized. Seven thousand people, he said, ‘“‘and they were acting much worse than you would at a high school football game.” A week later he was recruited by his friend Glenn Seaborg to chair the Department of Energy cold fusion committee.! Huizenga’s cochair on the panel was Norman Ramsey, a Harvard physicist who would win the Nobel Prize six months later. Ramsey, however, considered the assignment a “‘terrible job,” which left Huizenga with the bulk of the responsibility. Huizenga recruited the best scientists he knew for the cold fusion committee: in physics, for example, Dick Garwin, Steve Koonin, and Will Happer; in materials sciences, Howard Birnbaum; in electrochemis-

try, Al Bard and Mark Wrighton. He recruited Jacob Bigeleisen of the State University of New York at Stony Brook, because he thought no one knew more about tritium than Bigeleisen did, and Clayton Callis, president of the American Chemical Society, because he was foresightful enough to know that if the panel’s conclusion was negative, he would be accused of stacking the decks with disbelievers.* The directive given these experts was threefold: review the experiments and theory, identify experiments that should be done to test cold fusion, and suggest to DOE what its involvement should be. While in Santa Fe, Huizenga began arranging site visits to the various 303

304

BP GARY

TAUBES

laboratories that claimed positive results in cold fusion. Foremost among these was Pons’s lab. Pons and Fleischmann had skipped Santa Fe, but Jim Brophy was there, smiling and cordial. Brophy agreed to arrange a visit to Utah, but then he spoke with Pons, who argued that the committee was stacked with disbelievers. He would allow a site visit only if Huizenga added scientists to the committee who had already come out strongly for cold fusion. Huizenga refused. Meanwhile Pons threatened to have an injunction brought against the DOE panel, but Hugo Rossi managed to talk him out of it. When Huizenga returned to Rochester, he again tried to arrange the visit, arguing that it was in Pons’s best interest. Apparently Pons agreed, provided neither Wrighton, Koonin, nor Garwin were on the team that visited his laboratory. Huizenga acquiesced. Pons may not have wanted Wrighton because he still saw MIT as competition and didn’t want to give up his trade secrets. Koonin couldn’t make it anyway. (“If Pons really had something,’ Koonin said, “I should be the first person he’d want to come see it, because I’ll kiss his feet if he’s night.’’ Along these lines, Will Happer added, ‘“‘We’ll all carry him in a sedan chair to Stockholm. No problem.’’) On May 30 Al Bard spoke with Pons, who was still anxious that Huizenga’s panel was loaded against him. Bard suggested they add a couple of open-minded electrochemists: in particular, Larry Faulkner from the University of Illinois and Barry Miller of AT&T. Bard asked Pons if these two would be adequate, and Pons said they would help. So the final group going to Utah consisted of three electrochemists and two physicists: Bard, Faulkner, Miller, Ramsey, and Happer.? Bard also told Pons that the committee members would like to see all the data on a single operating cell—the calibration, history, everything. As Bard recalled it, “‘[Pons] said, ‘Fine, I have a cell ready. It’s getting excess heat right now; it’s putting out like mad, and you'll see it.’ ” The panel members arrived in Salt Lake City on the evening of June 1. Despite hoping to keep a low profile, they were met at the hotel by the local television news crews. The panel members declined interview requests, but that didn’t deter the press from following them from morning until evening the next day. Pons, after all his preliminary reservations, was as open as could be, with two exceptions: one, he did not have a cell putting out heat like mad, or putting out heat at all—he explained that there had been a power failure at the university. Two, he failed to provide a calibration curve, a crucial piece of data that Bard had specifically requested in advance.

BAD

SCIENCE

@ 305

The calibration curve would show the temperature rise in the cell as it varied with the amount of power put in through a heater.* Deriving a proper calibration curve is a long and tedious process, but without one it was impossible to know how much heat to expect if fusion was not occurring. Thus it was impossible to know what amount of heat was excess heat. That Pons professed not to have a calibration curve was virtually unbelievable. “Eventually,” said Happer, ‘“‘he said that maybe he’s got it at home. But here he’s expecting this high-level committee, and this is a key piece of data that he needs to analyze his results, and it just wasn’t there.”’ Pons did show the panel members his laboratory and his various fusion experiments. Now he had his cells set up in two Styrofoam coolers full of water, which were the temperature-controlled circulating baths. Happer had brought along a thermometer, and he moved it around the baths and found that the temperature was very stable. “So that was done right,” he thought. Happer was the kind of scientist who had to play with things with his own hands to understand them. Common sense was a major aspect of his experimental strategy. He was an atomic and molecular physicist at Princeton who did a lot of work with lasers. ‘I’m a tabletop experimentalist,’’ Happer explained. “That’s my training. That’s what I like to do. I like to do things that I can afford. In that sense, I probably have as good a background as anyone to go into a lab and tell what’s bullshit and what isn’t, because it’s similar to what I work with anyway.” Happer had decided upon hearing of cold fusion that it was probably wrong. In fact, a Scientific American reporter had called him a few days after the announcement, and Happer had harangued him for over an hour on the various aspects of fusion—its physics, the fatal effects of neutron radiation—that made cold fusion so implausible. ““The thing I didn’t have the nerve to do was say that just by looking at these guys on television, it was obvious that they were incompetent boobs.’’ Happer felt he was making no impression on the reporter, so finally he said that he would bet his house that Pons and Fleischmann were wrong. He believed in nuclear physics, and he knew “a little about the calorimetry, and what they were talking about sounded like it was close to the noise level.”” He assumed Pons and Fleischmann had botched the calorimetry, and the more he saw of cold fusion, the more he became convinced that

such was the case. “They just decided in their minds that it’s got to be right,” he said, ‘‘and they forced the data, forced themselves to make

everything support them.” Happer observed that Pons had calibration heaters in each of his cells,

306 # GARY

TAUBES

but seven of the heaters had never been connected to power supplies, which implied that the cells had never been calibrated. “They all had wires attached, coming up out of the cell,” Happer said. “‘And the wires were absolutely clean, the way the technician had stripped them. He was using alligator clips to attach them, so the moment he put the clip on, he would bend the wire. These wires weren’t even bent.” It was possible that Pons had just begun running those cells, in which case they might not have been calibrated yet. After all, Pons had said they’d just had a power failure, and Mark Anderson, who had taken over the cold fusion research after Pons fired Marvin Hawkins, later said they ran the cells for several days to reach equilibrium before calibrating. However, Pons also showed the panel members his new computer system and demonstrated how the computer was taking data on eight cells at once. It looked like a Potemkin laboratory. ““You have these things allegedly taking data,”’ Happer said, “‘but it was clear that only one of them could have ever been calibrated.” Pons did emphasize to his visitors that he had observed compelling bursts of heat, like the one he had reported in Los Angeles—5000 percent excess heat, for two days. These strange temperature increases were episodic and occurred without any rhyme or reason. Pons, of course, believed they must be fusion induced. “‘I think,’’ said Happer, “that Pons in his heart doesn’t believe that he can measure small amounts of heat. There’s no way his apparatus can do it, and it’s obvious that they’re careless about it. What convinces him, I’m sure, are these bursts. For that reason, they don’t care about the calibration.’ In fact, the

famous meltdown would have been aparticularly dramatic heat burst, and Pons and Fleischmann had based their initial faith on that episode. After the panel spent a few hours with Pons, Jim Brophy said that Utah also had its skeptics. One of the few was Milton Wadsworth, the dean of Mine and Earth Sciences. However, just a few days earlier Wadsworth, to his great surprise, had obtained positive results. So the entourage trooped across campus to see Wadsworth’s apparatus with the television crews trailing along behind. If Wadsworth was a skeptic, it was only by the rather low standards of the locale. He was an expert in a field called hydrometallurgy. He was also near retirement age and saw cold fusion as a way to make a big contribution in his last years. “We didn’t think for one second that there would be any difficulty in demonstrating the phenomenon,” he said, which doesn’t make Wadsworth sound like a skeptic. Back in April, Wadsworth had tried to obtain electrodes from Pons,

with the hope of dissecting and analyzing them. Pons had not been

BAD

SCIENCE

@ 307

forthcoming. Confidentiality agreement, he had said. So Wadsworth started running his own cells and found that Pons wasn’t giving advice either. Nonetheless, Wadsworth had done alittle electrochemistry over the years, and he had an able postdoc, an Indian researcher named

Sivaraman Guruswamy. The day before the Santa Fe workshop, Wadsworth told Brophy in confidence that they had observed two heat bursts. One lasted ninety-one minutes; one lasted twenty-four hours. Wadsworth’s heat bursts quickly made it into the local papers—‘‘leaky system,’’ said Wadsworth—and gave the administration faith that they were still on the night track. But Wadsworth was not a professional electrochemist, and neither was Guruswamy. Al Bard, for instance, noted that Wadsworth had had more

deuterium-oxygen recombination explosions than anybody else of whom Bard was aware. It was the mark of amateurs. As Larry Faulkner, who was Bard’s former student, explained, “‘One of the striking things in this whole field is that everybody was an amateur. Either you had physicists doing electrochemistry or electrochemists doing physics . . . and that led to a lot of strangeness.”’ Wadsworth showed the panel around his laboratory, which actually seemed in many ways superior to Pons’s. It was clear that Wadsworth had made an effort to calibrate his cells. Then Wadsworth showed the panel the data on his heat bursts. Here was one cell, for example, that went

along for eighteen days doing nothing at all. Then suddenly the temperature rocketed from twenty-seven degrees to over forty. The temperature remained high for almost two hours. Finally, they added heavy water to the cell and the temperature returned to normal. For one hundred minutes they had been generating the kind of excess heat that Pons and Fleischmann had advertised. Wadsworth said they had monitored both the temperature and the voltage during this episode. During the heat burst, the voltage going into the cell seemed to remain constant or very nearly so, which seemed like a good sign. They had an expensive, reliable power supply, so if both current and voltage remained constant, then the power entering the cell had remained constant. So what could have generated the excess heat? Fusion? What else? Listening to this, Will Happer began to feel like he had passed through the looking glass, and all the rules of experimental procedure had been rewritten. Maybe it was his vexatious common sense that was the problem. All the talk of fusion aside, the electrolyte in these cells was simply heavy water and lithium hydroxide, a lithium salt added to make the water conduct electricity.

308 BF GARY

TAUBES

Happer knew that these salt-water solutions are very temperature dependent—as their temperature changes, their resistance to electricity changes as well, and dramatically. In fact, if Wadsworth’s cells had really warmed from twenty-nine to forty degrees, the resistance should have dropped by half. Happer was familiar with these solutions from his youth in North Carolina. For several summers he had worked for a beekeeper,

and from this experience alone he knew that Wadsworth was wrong. ‘When you keep bees,” he explained, “‘you force the bees to build their cone in a hexagonal pattern that hangs in a wooden frame. But [bees}]wax is very soft, [so] to keep it from collapsing in the frame, you put wires in it to reinforce it. The way you get the wire to stay in is you heat the wire with electrical current until it just melts its way into the wax, and then you stop the current and the wax solidifies around it. Well, in Appalachia, transformers were expensive. So to get just the nght amount of heat, you used the current from your wall socket. But if you applied that to the wire, it would burn the wire out, because it’s too much current. So you have to limit the current with a variable resistor. The variable resistor was two electrodes stuck ina fruit jar with salt water. Now I’ve seen those things more than I’d like to say, and I knew perfectly well how they varied with temperature. As you would turn this thing on the salt water would get hotter and hotter, you’d get more and more current, and after a while the wire would start burning up. You had to keep cooling it, so I had a very good understanding of the temperature coefficient.” Happer brought this up while the panel members listened to Wadsworth, and the three electrochemists all immediately agreed. If the temperature in Wadsworth’s cells really rocketed upward, then the voltage should have simultaneously plummeted. After all, the resistance of the solution would drop, and it would take less voltage to pass the same current through it.° If the voltage hadn’t changed, neither had the temperature. If Wadsworth hadn’t been so anxious to find evidence of fusion, he

might have realized that he had a loose electrical connection that faked the computer into recording a temperature rise. Obviously, Happer explained, when

the researcher had added water to the cell, he had

inadvertently jiggled the wires and reestablished the connection. Everything was back to shipshape, and the temperature then appeared to drop. Happer and the three electrochemists suggested that Wadsworth make his electronics redundant. Thus, if one measuring device went haywire,

the other would still give the correct value. This is the kind of basic experimental caution that should have been routine. (Wadsworth and

BAD

SCIENCE

@ 309

Guruswamy made the necessary improvements and never saw another heat burst. In the meantime, however, the University of Utah had applied for a patent on Wadsworth’s work, and Pons and Rossi started talking up the importance of heat bursts rather than these steady-state heat measurements.) Happer found the entire episode mildly depressing. ‘““You know, instead of working this out themselves, they’re making this presentation to this distinguished panel in support of the Utah discovery.” The next morning the headline in The Deseret News read DOE PANEL SAYS U. RESEARCH MERITS FURTHER STUDY. JoAnn Jacobsen-Wells wrote that the panel “didn’t leave Utah on Friday professing to be card-carrying cold-fusion converts,” but they did conclude that Pons had an “‘intriguing phenomenon.” She then quoted Pons saying that the committee saw cells that were working, which was not true if he meant producing measurable and verifiable excess heat. ““They pointed out no area to me where there is a likely mistake,’”’ Pons said. ““That leads me to believe that they must believe what I have told them. It was a very positive meeting, and I think the interaction was excellent.” The New York Times noted that the committee was divided. It quoted one unnamed committee member saying they found no “show stoppers” and no “‘major errors” and no “smoking guns.” This source might have been Barry Miller or any of the three electrochemists, who all tended toward extreme civility when dealing with the press. Miller later said that since they hadn’t seen an experiment producing excess heat, they couldn’t prove the case one way or the other. “As far as we were concerned,” Miller said, “the experiment was still in limbo. We were

taking their word for it still.”’ When Happer saw The New York Times article, he was, in John Huizenga’s words, “‘firing mad,” and Huizenga had to talk him out of resigning from the committee. Happer felt that Pons had given them absolutely no reason to believe in cold fusion, which, in itself, was a

smoking gun. His indignation could have been seen as the product of a philosophical difference between physicists and chemists, except that he thought he knew who had been the Times’s unnamed source: Norm Ramsey. Not Miller or any of the electrochemists, but the soon-to-be Nobel laureate in physics from Harvard, the Eastern Elite school. Ramsey had seemed to be the most credulous of the panel members through the meetings. He had not, for instance, picked up on Wadsworth’s striking temperature-voltage problem. Once Happer explained it, Ramsey understood, but it took him longer to grasp the concept than it did the three electrochemists.

310

# GARY

TAUBES

As the panel members left for the Salt Lake City airport, Ramsey said the same thing to Happer that the anonymous source had said to The New York Times. Happer, who had a southerner’s fondness for storytelling, replied with a parable of a sort. “I said, ‘Norman, this reminds me ofthe old story of the rabbi. And his parishioner comes in and says, ““Rabbi, you know Iplanted this apple tree and it’s grown. I’ve watered it all these years and now it’s finally bearing fruit. But the biggest limb is hanging over my neighbor’s fence and my neighbor says it’s his apples on that limb. That can’t be right, Rabbi. I nurtured this thing.’’ And the rabbi says, ““Yes, yes, you’re right, my son, you’re right.”’ Well, not an hour

later the neighbor comes in: ‘Rabbi, you know this limb from my neighbor has been shading my garden all these years, giving me trouble, dropping leaves, and now it’s finally got some fruit. Surely, some of that fruit is mine, Rabbi, isn’t it? My neighbor says it’s all his.’” The rabbi says, “Yes, yes, you're right, my son, you're right.”’ And so the guy feels very good and off he goes. And then the rabbi’s wife, who’s been listening to all this says, “You silly old man, I never heard of any such thing. They can’t be both right.”’ And the Rabbi says, ““Yes, yes, you’re right.”’” Happer observed that after he told Ramsey the story, Ramsey replied with a smile, “Well, of course, yes, yes, you’re right.”’

2 CAMBRIDGE.

SALT

LAKE

CITY

When the results of Richard Petrasso’s lucubrations on the gamma ray spectrum appeared in Nature in mid-May, Pons and Fleischmann felt compelled to respond. They wrote a rebuttal, which Nature passed on to Petrasso for his response. Ron

Parker, Petrasso’s boss at MIT’s

Plasma

Physics Center,

had

hoped they could wash their hands of the affair after the first paper, but now they had to rebut the rebuttal. “This is crazy,” he said, adding that the Utah rebuttal was “‘absurd and pathetic” and to respond at all was to give it a dignity it didn’t deserve. Petrasso, by contrast, saw it as the opening in which he could now move in for the kill. “Thrust, parry, and counterthrust,”” he said. ‘“This time we’re merciless.”

The basis of Pons and Fleischmann’s rebuttal seemed to be that Petrasso’s original critique was meaningless because it was based on the

BAD

SCIENCE

@ 311

televised spectrum, the authenticity of which they could not affirm. Petrasso had attributed the authentication to Marvin Hawkins, but Pons

and Fleischmann rejected this. ““M.H.,” they wrote, “‘did not state that the quoted television spectrum was made in these laboratories, as it most certainly was not.”’ (To which Petrasso would reply, “‘It’s bullshit, is the kindest way to put it.”’) Pons and Fleischmann now offered their own gamma ray spectrum, which, curiously enough, appeared to be identical to the one that had been televised except it extended to higher energies, in a range of enormous peaks.

2.6146 MeV 20,000

Signal peak, 2.496 MeV

Counts

Compton edge of 2.6146 MeV

0

100

200

Channel number

These peaks, to Petrasso and anyone else who had ever worked regularly with gamma ray detectors, were instantly recognizable as the result of noise in the electronics. Indeed, Bob Hoffman said that he learned late

in the game that he had been working with a faulty preamplifier, and he told Marvin Hawkins to inform Pons. Hoffman would have liked to have seen an addendum to the rebuttal saying that an electronic glitch may have caused the gamma ray peak, but he wasn’t running the show.® Hoffman never told Petrasso about this faulty preamp, but he didn’t have to; it was obvious from looking at the spectrum. Petrasso knew that if the fusion pioneers had indeed induced fusion, their cells would have emitted neutrons with 2.45 MeV of energy. These neutrons would have then reacted with the hydrogen in the water bath, generating a 2.22 MeV gamma ray. With the entire spectrum in hand,

312 § GARY

TAUBES

Petrasso searched for a real gamma ray peak at 2.22 MeV, but it was clear that no such peak existed. Finally, Pons and Fleischmann had moved their peak in the rebuttal. Originally they had identified it at 2.22 MeV. Now they said it had an energy of 2.496 MeV. They didn’t bother to give an explanation for the move. Nor did they mention that no nuclear process existed that would emit a 2.496 MeV gamma ray. ‘How interesting,” Petrasso said. He also explained that the new extended spectrum allowed for the correct identification of the thallium 208 peak, which meant that what Pons and Fleischmann were calling their signal peak was actually at 2.8 MeV, rather than either 2.22 or 2.496 MeV.

This brought Petrasso back to the fateful question: why had Pons and Fleischmann labeled the peak originally at 2.22 MeV, when in reality it was at 2.8 MeV? (This was not answered by the fact that Hoffman had figured the peak at 2.5 MeV,

incorrectly as it turned out, and both

Hoffman and Hawkins had communicated this number to Pons.) When Fleischmann spoke at Harwell on March 28, he had reported the peak

at 2.5 MeV, and the Harwell physicists had replied, “It’s too narrow and in the wrong place.”’ The next day, Michael Salamon pointed out the discrepancy to Pons. As Salamon remembered it, Pons was already aware of the problem. “He was ready for it when we told him,”’ Salamon said. Pons replied that they had had acalibration problem, and in fact the peak was at 2.22, and he showed Salamon the errata he was about to send off

to the Journal of Electroanalytical Chemistry.? By the following Friday, when Fleischmann spoke at CERN, the peak was still too narrow, but

now in the right place. All this manipulation of the peak from 2.5 to 2.22 and back to 2.5, or 2.496, or whatever, is unfortunately what scientists classically consider fraud. The data were being modified to fit a preconceived conclusion: Pons and Fleischmann believed they had created fusion, they had a peak in the gamma spectrum, that peak should be at 2.22 MeV, so there it is. This explained Petrasso’s single-mindedness in pursuing the truth about the gamma rays. “Fraud is a pathology,” as Stephen Jay Gould wrote in The New York Times at about this time, although not on this subject. “I doubt that nonscientists realize how concerned all scientists are to purge any detected incident. The reason for our loathing is not widely understood, and its basis is not abstractly moral.” On June 6 Petrasso sent his latest rebuttal to Nature, along with an accompanying memo: “‘My coauthors and I,” he wrote, ‘‘feel we need to swiftly, forcefully, and publicly challenge (in print) this latest paper by

BAD

SCIENCE

@ 313

Fleischmann et al.”’ And then: “There is absolutely no discussion or justification of why they suddenly now identify their purported n-capture [gamma] peak at 2.5 MeV instead of at 2.22 MeV. Where did their 2.22 MeV identification come from? . . . Pons and Fleischmann cannot be allowed to escape answering this question. They must be directly challenged on this issue.”’ John Maddox, the Nature editor, took up the challenge to Pons and Fleischmann in an editorial that ran with the parry and counterthrust. However, he did not push too hard. The dispute over the spectrum had been “‘educative, to say the least,’ he wrote, and one detected the slight

scent of sarcasm. Maddox went on to say that none of what went on implied that the fusion pioneers had been “anything but straightforward,” simply that they had been under extraordinary pressure. “Put yourself in their position if you believe otherwise,” he wrote. Dave Lindley, an editor in Nature’s Washington office, said they didn’t want to be in the position of speculating on Pons and Fleischmann’s thinking process. ‘Using the F-word isn’t a good thing,” he said. “We felt in the end, the exchange between Petrasso and Pons and Fleischmann said everything there was to say.” It did seem obvious that, under the pressure of going public, Pons and Fleischmann had tried to do what they were hopelessly ill-equipped to do.* One could imagine the fusion pioneers in the frenetic rush, hoping for a peak, seeing one in Hoffman’s data, and not asking questions, simply assigning it the proper energy and moving on to the next order of business. It was not good science, but it was certainly human enough. ‘When the gamma ray peak business came up,” said Laura Garwin of Nature, “I was staggered by the incompetence of it. Fraud crossed my mind, as well. A good scientist doesn’t just say the peak is at one place and then finds out it is supposed to be someplace else and they say, ‘Oh, it is someplace else.’ It is equally explained by stupidity.” One chemist, who requested that he remain anonymous, proposed that the gamma ray business shouldn’t be considered fraud only because fraud suggests a rational motive. Without such a motive, one is left with incompetence or an insanity defense. ““They were caught up in it,” the chemist proposed, “‘and they just couldn’t analyze what was going on. They didn’t mean to defraud anyone. They just couldn’t assess right from wrong. That’s the most rational assessment of what went on.” Pons had sent Mike Salamon a copy of the rebuttal, and Salamon was appalled by what he read. He took it to Hugo Rossi and told him, mi Bhis is going to kill whatever credibility Stan has left in the physics community. It is very damaging. The spectrum is all wrong. He doesn’t address

314

BE GARY

TAUBES

the 2.2 to 2.5 energy shift. His interpretations are completely unjustifiable. People are going to look at this and laugh. It’s going to be major embarrassment.”’ Salamon then asked Rossi if he should tell all this to Pons, and Rossi said, “It won’t do any good, don’t even bother.” When Rossi saw the rebuttal, he felt ‘“‘almost as if some force had

wanted to play a dirty trick on us.”” And he added the obvious: “I think there was a lot of wishful thinking that went into the interpretation of that, which [Pons and Fleischmann] sincerely regret.”’

3 WASHINGTON,

D.C.

SALT

LAKE

CITY

When the cold fusion announcement was first made, Robert Bazell, the science reporter for NBC News, interviewed Robert Park, head of the

Washington office of the American Physical Society. He asked Park, off the record, whether he thought cold fusion was fraud, and Park said, No,

but give it two months and it will be. Park was proud of this piece of prognostication, even though by his own reckoning he was off by six weeks. He later said, “It became fraud the day [Pons and Fleischmann] received the helium analysis from Johnson Matthey and refused to release it’”’—that was June 6. (Although for Park’s prediction to be right, he also had to exclude from his definition of fraud the gamma ray shenanigans, the misrepresentation of the energyin/energy-out numbers, and all the various random prevarications and dissimilations that had come out of Utah.) At the Los Angeles meeting, Pons and Fleischmann had refused offers from MIT and Sandia to have their electrodes tested for helium because,

said Fleischmann, they had a prior commitment. A week later Jim Brophy revealed that they were committed to Johnson Matthey, the suppliers of the palladium, and that representatives from the company had been to Utah to collect the electrodes for analysis. (As for not taking up the offer made by MIT, Brophy said, “I can’t imagine any reasonable person turning [the electrodes] over to a competitor like MIT. That would drive our lawyers up the wall.”’) After hearing about the forthcoming Johnson Matthey analysis, Park began calling Brophy once a day asking about the status. Finally, Brophy said that scientists from Johnson Matthey were on their way to Utah, and

BAD

SCIENCE

@ 315

when Pons and Fleischmann had a chance to study their results, there would bea press conference. On June 6 Park called Brophy, who now said that Pons had decided not to reveal the results of the helium analysis because it would violate peer review. Park put this in his weekly newsletter, and that was the last mention of the Johnson Matthey helium analysis. (A few weeks later, Rossi reported confidentially to the Utah cold fusion committee that General Electric had tested Pons’s rods for helium and found none, but

that was a different analysis.)

4 PROVO

On June 13 another subset of the Department of Energy panel visited BYU, and Steve Jones gave them the grand tour of his facilities. Afterward, Steve Koonin, who was among the group, told a reporter from The Deseret News that he was still not convinced that anyone, in particular Jones, had concocted cold nuclear fusion, although he was “leaning toward that direction.’’ Koonin, however, was a theorist, and he had a

theorist’s naiveté when it came to experimental details. He was also attempting to be diplomatic and undo some of the political damage he had caused in Baltimore when he accused Pons and Fleischmann of ‘incompetence and perhaps delusion.”’ John Schiffer, another panel member, came away from the BYU visit still unconvinced that any nuclear phenomena had occurred in test tubes or baby food jars. Schiffer was an experimental nuclear physicist with thirty years’ experience. He held a joint position at the University of Chicago and Argonne National Laboratory and had recently been elected to the National Academy of Sciences. John Huizenga, for one, believed that, as a generalist in nuclear experiment and theory, Schiffer was ‘‘probably one of the best.’’?° Schiffer quizzed Jones on the details of his statistical analysis and came away thinking that he was a likable guy but not the most careful physicist he had ever met. Jones didn’t seem to understand what Schiffer called “simple statistical methods for treating data.’ He later remarked that Jones ‘‘wasn’t terribly interested in digging in and establishing whether there was an effect or not.”’

316 @ GARY

TAUBES

On June 15, Harwell finally went public. For two months Dave Williams and Derek Craston had been sixteen cold fusion cells, designed from the information Martin mann had provided. They had four sizes of palladium electrodes electrolytes: four with lithium, two in heavy water and two

running Fleischand four in light

water; four with sodium, two in heavy water and two in light water.

They had a heater in each cell to calibrate it; they measured the voltage across the heaters, the current going into the heaters, and the current

going into the electrodes and coming out of the electrodes. They knew exactly how much power they dissipated in electrolysis, and their computer logged it all automatically. The computer system itself had taken three weeks to set up and debug. They had separated the sixteen cells with the controls in one water bath and the test cells in the other. They had neutron detectors mounted over each bath, so if more neutrons were

coming out of the test cells than the controls, they would know it. Williams had promised Fleischmann that if there was an effect, he was going to make it happen. “I thought at the time,”’ Williams said, “I might just try and make something blow up if I have to. I am going to bloody well nail this. Martin knew that I really wanted to prove him right.”’ Instead, Williams and his colleagues discovered several critical sources of error in the Utah calorimetry.’? Had Fleischmann been aware of these when they began, said Williams simply, ‘“he would’ve told us.’’ Some of the Harwell cells did seem to run excess heat for just these reasons. “The problem,” said Williams, “was the ones that seemed the best were ones that had ordinary water in them.” This was yet another lesson in running controls. Williams added that if they averaged out all the results, the heavy water cells ran slightly hotter, but less than the error in the measurements, which meant it was still compatible with zero. Nonetheless, he said, “if you were an evangelist for cold fusion, you could say, ‘Hey, these guys have demonstrated cold fusion.’ ’’ When they ran the cells in a foolproof calorimeter, this artifact vanished. As of June 15, Harwell had detected no neutrons and no anomalous heat. They had prepared electrodes a la Huggins and seen nothing. They had run cells by Bockris’s prescription and seen nothing. They had run titanium chips in deuterium gas as Frascati had done and Jones was now doing and seen nothing. They studied the enhancement of tritium in the cells, and it only confirmed what had been in the literature all along.

BAD

SCIENCE

@ 317

Williams had written to Fleischmann saying they’d seen nothing and would have to pack it up soon, and Fleischmann called him up all cheery, and said they were still getting positive results in Utah. The Harwell cold fusion crew had even signed their own confidentiality agreement with Johnson Matthey, and the metal company had loaned Harwell palladium electrodes. ““They said to me,’’ Williams recalled, “‘we will supply you with samples of all the types of palladium we supply. So I got the lot.”’ As for whether these rods generated anything of interest, said Williams, ‘‘Zilch.”’

By June 15, the Harwell cold fusion effort had cost one third of a million pounds, about $500,000. Williams now told reporters that all this

had come to nothing. ‘““What we have to recognize,” he said, “‘is that brilliant people have mad ideas.’’" Many of those who already did not believe in cold fusion considered Williams’s pronouncement

the final word.

Fleischmann had, after all,

given Williams and the Harwell researchers detailed instructions on how to build and run the cells, hadn’t he? Al Bard later met Fleischmann at

a conference in Stockholm, and Fleischmann said he never gave them direction at all; he only sat down over lunch with them once and talked it out. He also said that the Harwell people never believed it, and that if you really don’t believe something deeply enough before you do an experiment, you will never get it to work. Bard observed that Fleischmann had seemed very unhappy talking about Harwell. The folks at Johnson Matthey also rejected the Harwell work. They told Rossi that Harwell made this announcement because the United Kingdom was privatizing its fission industry, and positive cold fusion results would drive the price down. At this time Johnson Matthey was negotiating with the University of Utah to let the administration in on the secret of their magic palladium. In exchange for the recipe, Johnson Matthey was asking only 40 percent of the potential cold fusion royalties. (Said Hugo Rossi, ‘‘And they say Utah is the schlock capital of the world. Such nerve!”’) The logic at Utah, as Rossi recalled it, was that “Johnson Matthey knows how to make the right stuff, nobody else does,” which contradicted Huggins-Stanford, who said he used off-the-rack palladium and got positive results. Utah eventually passed on this opportunity, although Chase Peterson went so far as to contemplate the pros and cons, which is to say they’d be trading away either 40 percent of zero or 40 percent of infinity. So what did they have to lose? When the Los Angeles Times asked Jim Brophy for a comment on the Harwell finale, he replied, “It is disappointing that they have not been able to do the experiment properly.” Bockris later insisted that Williams

318 # GARY

TAUBES

had indeed confirmed cold fusion because he had seen “‘silly little bursts” of heat in his cells. He ignored the fact that Williams had seen silly little bursts in cells run with tap water. He agreed with Brophy here. As far as he was concerned, Williams didn’t know how to do it right. “That 1s

the whole point of the phenomenon,” he would say, “‘you have to know how to do it and then maybe you get it.”

6 COLLEGE

STATION

On June 16, the Department of Energy panel visited Texas A&M. John Bockris distributed a set of guidelines to the panel members so that there would be no misunderstandings about the A&M cold fusion work. The six-page memo read like something the Queen of Hearts might have written had she left Wonderland and gone into nuclear research. Bockris summed up the important points in a few paragraphs: It has always been the anomalies which can be seen in a Science which gives rise to the new ways of thinking which cyclically invade the sciences. The constant reiteration of the old way (particularly with the great Anger and Emotion)"* we are seeing among our colleagues and visitors has not been the way that changes in scientific attitudes have come in the past. Therefore, when persons tells us that they have carried out the electrolysis of deuterium evolution in palladium and see nothing new, particularly if (as is usual) they are furious about it, have spent little time on it, and have

little experience as to how to do experiments of the type named, we tend to discount their contribution.

This description could be considered the modus operandi of the ebbing field of cold fusion. And then came Bockris’s prescription for future research: “At the time of writing, the phenomenon is less than three months old. Two or three years (5-6 centers, 100 people) will be the right sort of time to think of in order to make a decision as to whether it is worth Big Money.”’ This was a variation on Ira Magaziner’s “‘highfalutin” thesis. Give us small money now and we will try to do good science, and let you know if it was worth it. (Bockris would later explain that, of course, his research was sloppy, but that was because it was funded in “drips and drabs.”’ Once he received proper funding, he would do it better. It would go ‘‘zoom.’’)

BAD

SCIENCE

#@ 319

Bocknis expounded on this theory: “Our attitude is that we may be in an emerging area of science, and that in such situations experiment usually molds theory to fit it... . We are particularly unenthusiastic in the discussion of the application of present theories of fusion in plasmas to the idea of fusion in electrochemical confinement.”’ In other words,

don’t give us any of your nuclear physics. Bockris had already formulated his own model to explain the experimental observations of cold fusion. This was his dendrite model, which

went, in brief: First, electrolysis induced the growth of dendritic spikes on the electrodes, what Bocknis called “spiky protrusions of the depositing metal.”’ Then, ‘‘abnormally high”’ electric fields fired up at the tips of these dendrites. These fields generated electric sparks.'® The sparks ionized the deuterium, which then was whipped by the intense fields against the electrode with an energy of 1000 to 2000 electron volts. This energy was enough to cause an occasional deuterium nucleus to crash the repulsive barrier of another deuterium nucleus and fuse. Voila. The result was a kind of hot fusion, but fusion nonetheless.

Bockris said that, by twiddling a parameter here and adjusting a variable there, he could ‘‘rationalize,”’ among other data, the levels of tritium

he had observed. He called it “an adventitious, funny phenomenon, which turns on in strange conditions.”’ If nothing else, he said, “the facts

are certainly correct.” No less an authority than Stan Pons himself had told reporters that Bockris’s dendritic model might be relevant to his results. After Bockris explained this theory to the DOE panel, Dick Garwin told him that the one thing he knew for certain was that if ten volts are applied to the cell, no matter how one imagines a deuterium nucleus scooting around inside, it is never going to end up with more than ten volts of energy. This is high school physics. It was a question of conservation of energy, if nothing else. “You are kidding yourself,” Garwin said. Bockris replied that they had to put aside all understanding of these things until the scientists involved demonstrated definitive results. When the panel broke for lunch, Norm Ramsey was assigned to explain basic physics to Bockris. While the others ate, Ramsey was up at the board writing simple equations: the integral of the electric charge over time, and so on, and how that relates to potential difference. Bockns

either could not or would not understand. “He was just absolutely certain that his own theory was right,” Will Happer recalled. “And he had this collection of pathetic students and postdocs all applauding him when he spoke. It was very sad.” (John Schiffer later observed that Bockris didn’t grasp any of the principles of physics but enjoyed using the

320 B GARY

TAUBES

words. Kevin Wolf later said that he also tried to explain the reality of basic physics to Bockris and failed. And Roger Parsons, Bocknris’s former student and the editor of JEAC, said he tried to talk Bockris out of publishing his dendrite theory and also failed. “If people really want to publish,” he said, “then I let them.”)

The panel spent the morning listening to the Texas A&M researchers present their data. John Appleby, whose excess heat data, along with Bockris’s tritium, was later credited with keeping the field alive, spoke for forty-five minutes. Even Kevin Wolf acknowledged that he bombed. Appleby himself would admit that he was not even a desk scientist so much as a manager and a fund-raiser. For cold fusion, Appleby’s second in command, Supramaniam Srinivasan, was the desk scientist, and Young

Kim, a Korean postdoc, had done the great majority of the lab work. This was the third time, at least, that Appleby had presented his cold fusion data, but the first time he had to vigorously defend it, which he could not do. ‘Appleby didn’t know anything about it,” Wolf recalled. ‘He would say things that were just wrong. Finally it got so bad that [Srinivasan] corrected him. And Srini even tripped up and corrected him incorrectly once and had to take it back. Oh my God, that was a fiasco.” At one point Schiffer asked Appleby to show them the raw data from a huge episode of excess heat that he had displayed on his transparencies. The data were originally taken offa strip chart, so Appleby sent Kim back to get it. When Kim returned, it turned out that this dramatic excess heat event was in reality a blip on the chart, barely visible. Indeed Happer said he had to strain his eyes to see it, and Clayton Callis remarked that the point of the pen used to draw the line was wider than the blip itself.'¢ Chuck Martin was present during this talk, where he also saw Appleby’s raw data for the first time.'” He said he was “‘flabbergasted’’ when he saw the infinitesimal deviation that they had displayed at two conferences as undeniable excess heat. It got worse. Appleby presented this blip on the strip chart as though his equipment had remarkable accuracy. In scientific lingo, Appleby presented the data to at least three significant digits. It was as though he had looked at a glass of water and rather than saying it was half full or half empty he had reported that it was .523 full. The panel members asked Appleby how he could get such incredible detail from reading a squiggly line on a strip chart. Schiffer recalled that his reply was ““‘We had graduate students with good eyesight.”’ It got still worse. What Appleby represented as excess heat often turned out to be cases in which the temperature had remained, as Schiffer put it, “rock steady.’’ Instead his equipment had recorded a minute dip in the recorded voltage.'* Appleby

BAD

SCIENCE

@ 321

deduced from this that since the input power had decreased, and the temperature had remained more or less constant, he had witnessed excess heat that could be caused by fusion. “I couldn’t believe it,” said Martin.

“It was rather straining the imagination in the way they were presenting it,” said Barry Miller, who was also an electrochemist. And Will Happer, a physicist, said, “If you actually looked at the raw data, you could just barely see the needle moving on the chart paper. That plus the fact that they had no understanding of their instrument, and it was consistent with zero effect. Anybody without any preconceptions would have said there’s nothing there.” Later Appleby told a reporter that the DOE panel had simply misunderstood how they read their instruments. This may have been true, except then Wolf, Martin, and several graduate students had misunderstood as well.’® Bockris also presented his experimental data, which demonstrated 5 to 10 percent excess heat. Miller, however, noted that Bockns was only

measuring a one-degree rise in temperature, and he was claiming this was excess heat. Bocknis’s graphs had no error bars, which would simply have indicated the possible error in each measurement. The absence of error bars seemed to imply that the measurements had no error. This was unusual in science because the experimental measurement has not been invented that doesn’t come with some measurement error. Calculating and presenting the limits of error in an experiment is usually taught in high school science classes. As E. Bright Wilson put it bluntly in An Introduction to Scientific Research, “A measurement whose accuracy is completely unknown has no use whatever. It is therefore necessary to know how to estimate the reliability of experimental data and how to convey this information to others.” When the panel members queried Bockris on measurement errors, he replied, absurdly, that he avoided them by personally training many students to take the measurements. ‘“We used to alternate the students,”’

as Bockris put it, “‘so we got objective measurements of the calibration on the heat cell. If Jack does it like Jill, it is bound to be nght.” After the morning presentations, the DOE panel members split up and visited the different A&M labs. Happer, Schiffer, and Miller were among those who went to Bockris’s lab. Happer proceeded with his usual poking and prying. He later said that Bockris’s calorimetry resembled a Rube Goldberg contraption, which was not unkind if one considers that Nigel Packham, Bockris’s graduate student, had described it as “primitive as hell.”

322 BP GARY

TAUBES

When Happer inspected the apparatus closely, he found that all the wires were very loosely connected with clamps and alligator clips. He noticed, as did Schiffer, that if he touched one of the alligator clips, the

needle on the chart paper recording the cell temperature jumped up a few degrees. If he did so again, the needle jumped down a few degrees. So a simple touch could induce the calorimeter to record an effect equivalent to what Bockris had reported as excess heat. “Tt was clear to me,’’ Happer said, “that I could get any answer I wanted by the accident of howIleft [the wires and connections] the last time I touched [them]. So we asked [Omo Velev, who was doing the calorimetry], ‘Could you show us how you put additional water in here?’ So he gets the syringe, and he sticks the water in, and he could not do it without brushing the wires. Sure enough, as soon as he took the syringe out there was a jump on the chart paper, about the same order of magnitude of the effect they were seeing.” Happer asked Velev how they could ever measure heat to any accuracy, because obviously the error in the apparatus was at least 5 to 10 percent. Happer suggested that they must see negative results as well at this level. Velev replied that negative results happened all the time. Let me show you, he said. Velev pulled open a drawer and showed Happer, Schiffer, and Miller page after page of chart paper with negative results.?° But Bockris had said that he wasn’t interested in negative results: “Negative results can be obtained without skill and experience.”’ Apparently, this philosophy went for his own lab as well. When his cells appeared to produce less heat than they should, these were considered mistakes and stashed in a drawer. When the cells appeared to produce excess heat, this was considered evidence of fusion. In a conventional

scientific endeavor, this would be called cooking or trimming the data. Cold fusion, of course, was not a conventional scientific endeavor.

Finally, there were the tritium observations, which had become the most conspicuous, if not the only, evidence for cold fusion. Bockris and Packham, who had done the work, now reported that they had observed

massive doses of tritium in nine cells. Wolf, however, reported that he had seen no neutrons of any consequence emerge from either of the two Bocknis cells that showed tritium while in his laboratory. A neutron detector had been monitoring at least one of those cells, but it had not even registered a peep when the tritium appeared. This troubled Jacob Bigeleisen, the tritium expert on the panel. He was another veteran of the Manhattan Project and nuclear weapons programs. He’d studied the chemistry of tritium, the enrichment of tritium, the kinetics of tritium, the vapor pressures of tritium com-

BAD

SCIENCE

@ 323

pounds. At Los Alamos back in 1955, he had created a technique, now used widely, which removes tritium from the heavy water used in nuclear reactors. Regardless of what Bigeleisen thought about cold fusion, there was not much he didn’t know about tritium.

Bigeleisen asked Bockris, Wolf, and Packham why they believed they were creating tritium without “balancing the equation.” This was high school physics again—the conservation of energy and momentum. If two deuterium nuclei fuse, what comes out of that reaction must surely have the same amount of momentum and energy as what went in. Bockris could not just say deuterium plus deuterium equals tritium, because the energies and masses on both sides of the equation would not balance. It would be like saying two plus two equals three. “They just can’t be making tritium and nothing else,’’ Bigeleisen said. ““They have to show what else comes about. They have to give a material account. They can’t just say D plus D equals T.” To do so is not science. Everything known about nuclear physics up until March 23 dictated that if fusion had generated Bockris’s tritium, then 10 billion neutrons per second should have been generated along with it.”* That is the other side of the equation. Bigeleisen later said he solemnly warned Bockris that if he truly believed he had generated tritium by fusion, he had better watch out because the associated neutrons would constitute a serious,

even fatal, health hazard. Wolf's detectors had registered one neutron per second, so he was off byafactor of 10 billion. But, as Bockris had made

clear, this was orthodox nuclear physics. Bigeleisen added that if they had generated tritium, they had to have generated a proton to balance the equation: D + D > T + P. Four nucleons on each side of the equation. That proton would strike the palladium in the lattice, and the palladium would emit a gamma ray of a specific energy. This was a well-studied nuclear reaction, which has nothing to do with fusion. Wolf's equipment hadn’t detected any gamma

rays. Then, asked Bigeleisen, where were the secondary neutrons? The fusion of two deuterons creates a tritium nucleus with tremendous kinetic energy, which is to say it’s spit out of the reaction as though fired from a cannon. Physicists had learned over the years from working with plasmas, atom smashers, and hydrogen bombs that these [tritium nuclei] fuse with deuterium very favorably and generate fourteen MeV neutrons. Cold fusion cells brimmed with deuterium. Schiffer had worked out the calculation with Garwin:

10,000 to 100,000 of these neutrons

should have emerged each second from Bockris’s cells. Wolf had not observed these neutrons either, but he admitted that he should have.

324 BE GARY

TAUBES

Had Bockris been less willing to disavow the entire body of knowledge on nuclear physics, he might have considered that his tritium hadn’t been generated in the cell by cold fusion, but rather had entered the cell in some other manner. The odds that the tritium had actually been generated in the cell by fusion seemed virtually astronomical. A billion to one, maybe? Ten billion to one? Who could say? But Bockris professed to be convinced. And Packham presented evidence that he said demonstrated the production of tritium over time in the cell. This was the data from the cell, known as A7, which Packham and his colleagues

had sampled for over twelve hours back on Apnil 28. Packham showed the following curve:

800 ek 600 €

E 4 ~

Qa

=

400

55

=

S

200

0 0

2

4

6

8

time (at high i) /h

10

12

14

BAD

SCIENCE

@ 325

Bigeleisen took a look at the curve drawn on four points and said, “Well, your data do not uniquely define that curve.’’ He suggested they could equally well draw the following curve through the same four data points: 800

=

600

£ £

aea >

=

400

2

G

5

oO =

200

:

| 0

2

4

6

8

10

12

14

time (at high i) / h

Then Wolf asked Bigeleisen, ‘Jake, are you implying that someone spiked that sample?”’ And he replied, “Kevin, you said that. I would never say such a thing.” He added, “You know people say the heat appears in bursts; the neutrons appear in bursts, and maybe the tritium appears in bursts.” As Bard remembered it, Bigeleisen suggested, “You know, there’s another explanation for a curve that looks like this.” Either way, the implication was obvious. Later, when they toured the labs, Bigeleisen asked Packham whether they had a supply of tritiated water in the building. (Tritiated water is similar to heavy water but contains tritium rather than deuterium.) Bigeleisen knew that if the cells had been spiked with tritiated water, the result would look identical to what Packham had reported. Tritium is also a tenacious contaminant, and it could get into the air, into the water,

even into the stopcock grease, and eventually be reported as a confirma-

tion of cold fusion. Said Bigeleisen, “You have to be very careful with

tritium to prevent contamination.”””? Packham told Bigeleisen that they did have a bottle for radioisotope on studies, but it was on the third floor, and the cold fusion cells were

326 B GARY

TAUBES

the first. “Could that get contaminated down here?”’ Bigeleisen asked. And Packham said no, it was under lock and key. Ramesh Kainthla, Bockris’s chief researcher, later said that was not the case, and Packham later admitted that the bottle of tritiated water had not

been locked up after all. Still, he said he had been secretly checking on its level and activity, and neither had changed. He also observed that the tritiated water they had in the lab was plenty potent enough to account for all the tritium in their cells. ““That doesn’t belie any fears of people spiking samples,” he said, “‘but it’s just ridiculous to think that anyone would.” Packham seemed nervous even talking about the tritium. If the cells had been spiked, he was the prime suspect. Bockris himself would say, “This story of Nigel putting [tritium] in the solution is all over the place. Everyone thinks he put it in the solution. That has been with us since the beginning.” Even if the cells hadn’t been spiked, it seemed that there was something monumentally wrong with the work. That the subject made Packham nervous, of course, made him seem all that much more suspicious.

The entire edifice of cold fusion was perched precariously on the observation of the tritium, and it all rested squarely on Packham’s shoulders. No less a celebrated crowd than Hugo Rossi; Chase Peterson; Jim

Brophy; Milt Wadsworth; Robert Hoffman, who did Pons’s gamma ray work; Marvin Hawkins; the funding agents of the Electric Power Research Institute, who were providing support for Bockris; Steve CrouchBaker and Turgut Giir, who did the calorimetry experiments for Bob Huggins; Norman Ramsey; and at least once, even Stan Pons said their willingness to believe in cold fusion came from A&M tritium.2? As Rod Decker, a political reporter in Salt Lake City, said about Pons, ‘This tritium has given him the serene confidence of a Christian gentleman holding four aces.”’ Packham was a small, slender man. His hair was long by the standards of College Station, Texas, and he wore a small gold earring in his left ear. When he took on cold fusion, Packham was already in his fifth year with Bockris and his eighth year of postgraduate work. He did an undergraduate degree at London Polytechnic, then went on to Imperial College, where he spent three years doing postgraduate work in spectroscopy. Packham would say his Ph.D. was still pending on that, but his adviser, Alastair Gebbie, said Packham had only been a laboratory assistant, not on any degree track. Packham had served him well, but he was relieved when he moved on. Gebbie said he couldn’t in good faith sponsor Packham for a higher degree and didn’t feel he was cut out to be a

BAD

SCIENCE

§@ 327

physicist. Mino Green, a chemist at Imperial who had studied with John Bockris, got a call from him one day when Bockris was scouting for people to join him at A&M. Said Green, “‘Nigel was looking to go to the States, so I put the two together. . . . [He was] an enthusiastic young man, resilient enough to work with John.” Before cold fusion, Packham had been working on the bacterial pro-

duction of hydrogen, a biochemistry problem that would have been his thesis. There’s some question about how far along he was, even after five years. Packham said he could have written up his research and been gone by Christmas, but he’d been sidetracked. Now the aborted bacteria work

was only going to constitute half his thesis; the generation of tritium in an electrolytic cell would be the other half. This seemed like a dubious proposition. What if the tritium turned out to be contamination, or worse? Could a graduate student write a thesis on a botched experiment? How about half a thesis? Jeff Wass noted that it was Bockris who wanted it more

than Packham,

that Bockris felt

awkward making Packham do six months of work on cold fusion and putting off his thesis work by at least that long. ““So they made some type of reconciliation, primarily on Bockris’s own decision, that this would be

part of his Ph.D.” Packham later said he was doing it because it was an important part of his time at A&M, and that although it only represented a year of his life, it was so intense that it was “probably worth about two or three years of normal research.” The suspicion that Packham’s cells had been doctored with a smidgen of tritium had been circulating more and more widely since Santa Fe. Some chemists simply refused to believe that John Bocknis, of all people, would be the first chemist to concoct a magic recipe that transmuted deuterium into tntium.”° Frank Cheng, who was Chuck Martin’s postdoc, started wondering if the tritium was legitimate as early as the middle of May. Cheng was aware of the quality of results that Bockris habitually published. He said, “T don’t know if someone in that lab was going out of their way to please Bockris or not.” By the end of May, Cheng had heard that Bockris had bottles of tritiated water in the laboratory. Omo Velev even came up to him one day in the hallway and said that he could see how some of this tritium could migrate over to the fusion cells by accident or sloppy lab work, but nobody seemed to be doing anything about it. By the time the DOE panel arrived, Del Lawson and Cheng had constructed and run nearly fifty cells, hoping to reproduce in one of them, at least, Bockris’s tritium results. They tried every imaginable configuration; they used electrodes of all different sizes; they added

328 BH GARY

TAUBES

poisoning agents like hydrogen disulfide, bromide, or iodide. They found, as Martin put it, “nothing, nothing, nothing.”

7 SALTLAKE

CITY.

GROUND

ZERO.

PHYSICISTS

For five weeks, Mike Salamon had been in the lion’s den, or ground

zero, depending on your choice of metaphor. He had run his gamma ray detector in Stan Pons’s basement laboratory for 831 hours before finally removing it on June 16. In those weeks, Salamon had been informed only once of the existence of an active fusion cell: Mark Anderson called one night saying he better come over because they were boiling water. ‘““We have an active cell,’’ Anderson said. (Because Pons was doing all the calorimetric analysis himself, the only way Anderson knewacell was “‘active’’ was if it was

boiling, or perhaps melting down through the concrete heading for China.)

Salamon

ran over to Pons’s laboratory,

and there Anderson

pointed out the boiling cell steam spewing from the top. Salamon had his gamma ray detector running under the table, monitoring the episode, but he wanted to observe neutrons directly as well. He told Anderson not to touch the cell. He was going to the nuclear engineering department to make up aset of neutron activation foils and he’d be back. When Salamon returned two hours later, he was horrified to discover

that the cell had been turned off. Anderson said Pons had told him to turn it off so Salamon called Pons, begging him to turn it back on. Pons argued that Anderson couldn’t hang around all night just so Salamon could measure neutrons. Besides, Pons said, it was very expensive to boil

off heavy water like that; it had to be replenished continually. There was nothing Salamon could do. It was Pons’s lab. Pons later told him that they had obtained tritium from this cell, which made it all the more

bittersweet. For the five weeks that Salamon’s equipment had run in his lab, Salamon rarely saw Pons. He was hardly ever in the lab in the afternoons, which prompted Salamon and others on campus to speculate that he had a hidden laboratory somewhere; quite a few of the Utah faculty began to suppose that Pons had simply moved the entire project to his house. John Gladysz, for instance, who had collaborated occasionally with Pons,

BAD

SCIENCE

# 329

said that as time went on, he kept readjusting the odds against cold fusion. As they got longer and longer, he was left with a single scenario that allowed for cold fusion to be possible, although not probable. ‘‘Stan is a pretty bright guy,” Gladysz said, ‘‘and knowing what has pissed him off in the past . . . I would not be surprised if Stan just moved the whole project off campus to his house.” That would explain, among other things, why the cells in Pons’s lab never seemed to be working. Under this scenario, Pons’s modus operandi would be to assure that nobody would duplicate cold fusion and become a commercial competitor, because nobody would ever see what went into areal cell. Ergo, he

was even suspicious of his own university, the administration of which he certainly blamed for some of his woes. Hugo Rossi later said Pons spent his time at home analyzing the data from the laboratory, listening to the rock group Dire Straits, and drinking red wine. Pons had told Rossi that his rule was “no experiments at home.’”’ And as Nate Lewis later observed, the theory didn’t work “because they’re bringing along Huggins and Bockris and everyone else on their coattails. If they really want no one else to know, why would they think those other guys can do it? It does not make sense.” In any case, Salamon pulled out of the lion’s den when Pons left for England in June. Pons told him he had purchased a new detector and needed the space occupied by Salamon’s equipment. Even Salamon admitted that the new detector was exactly what Pons would need for detecting low-energy neutrons. But, as Salamon said, “our cause wasn’t helped by the fact that we had no evidence of fusion.”

8 SALT

LAKE

CITY.

ADMINISTRATORS

They were losing the propaganda war, wrote Hugo Rossi on June 19 in a letter to the two fusion pioneers, and the fact that he prefaced this remark by saying, “I really do not want to make this letter intolerably unpleasant, but... ,” strongly suggests that even his alliance with Pons and Fleischmann was deteriorating at this point. Rossi’s letter seems to have been a diplomatic attempt to convince the two chemists that open cooperation would be more productive than the “variety of obstructive schemes” they had employed so far. Rossi was

330 § GARY

TAUBES

fatalistic. He insisted that their only hope was to do good science while simultaneously predicting that the eventual demise of cold fusion had already been decided: ““We have to lose, for there is no way to do science other than at the pleasure of the ‘Masters of the Universe.’ ”’ He concluded:

There was a cartoon in the paper a few days ago depicting the Ayatollah approaching his final residence in hell. He was being directed by the devil to begin reading the literature which was waiting for him: thousands and thousands of copies of The Satanic Verses. Instantly the image of myself in the same predicament shot up: I was being required to read thousands of copies of the paper of Nate Lewis. Watch out! We’re all going to hell together.

So Rossi was off to purgatory, but not before trying to force reputable science out of Pons and Fleischmann. In the same letter, Rossi sent Pons and Fleischmann the operational plan for a second helium analysis, an idea Fleischmann had originated after Los Angeles. Both Sandia and MIT had offered to analyze the Utah rods for helium, but Pons and Fleischmann perceived the offers as arrogant. As Rossi told it, “These people came on saying that any lab with a ruler in it could find the kind of helium that was necessary for this helium 4 theory to make sense, and they said send an electrode, and in few days we'll give results. This was a gross oversimplification, and it was Fleischmann’s idea to call them on it.” The procedure was as follows: Pons would take five palladium electrodes and send them to the mediator, John Morrey, a physical chemist at Pacific Northwest Laboratories (PNL) in Richland, Washington.?6 Morrey’s researchers would slice each of the electrodes into pieces and distribute the pieces to six different laboratories. Each of these would in turn analyze the electrode segments for helium and send the results back to Morrey. Pons would not provide the history of the rods until after the analysis was complete, so the researchers would not know which rods had produced excess heat. And Pons would have no knowledge of the analysis. It was, in the lingo, double blind. Rossi, however, later took to referring to the analysis as the ‘‘double-

blind double-cross.” It was probably this episode that led Rossi to toy with the notion that the entire cold fusion affair was a scam, orchestrated

by Fleischmann himself from its inception. Fleischmann’s motive, in Rossi’s scenario, would be to render the entire scientific world,

particular the stuffy, self-righteous gang of physicists,” ridiculous.

“‘in

BAD 9LOS

SCIENCE

@ 331

ALAMOS

When cold fusion cells at Los Alamos generated tritium on June 20, the news spread instantaneously. Apparently, Ed Storms, who was doing the work with his partner Carol Talcott, told John Bockris, and immediately the information reached JoAnn Jacobsen-Wells at The Deseret News. The June 23 News ran a story headlined N.M. LAB REPORTS FUSION BYPRODUCT. The story represented the first time the Los Alamos administration had heard of these results. Since Los Alamos had a well-established procedure for dealing with the press, the administration was, as a press officer put it, ‘‘heartbroken.’’ Storms later said that he had his feet ‘“‘held over the

coals” by management for his little indiscretion. The Los Alamos administration suspected that the results had been leaked to help prop up the cold fusion efforts in Utah, which were scheduled for confirmation hearings on June 29. It’s equally possible that Bockris leaked the story because he considered the Los Alamos tritium confirmation of his own results and evidence that his results were not bogus. Ed Storms was a chemist by trade who had been at Los Alamos for thirty-one years and worked in a variety of programs, including a nuclear rocket project known as Rover and some high-temperature superconductivity studies. Storms had an impressive demeanor and a messianic glow in his eyes, like Charlton Heston playing Moses. His partner in cold fusion, Carol Talcott, was thirty-two. She had followed her husband, a

physicist, to Los Alamos five years earlier. She had been working toward her master’s degree in atmospheric chemistry but never finished it. At Los Alamos she worked with palladium hydrides, which led her to look at cold fusion as a natural progression in her research. Storms and Talcott said they believed cold fusion was correct as soon as they heard about it. Storms said the simplicity of it just hit him. “And

even if it weren’t correct,”’ he said, “‘it would still be an exciting area to

be into.” Bockris might have said they were a perfect pair to study cold fusion: they were not biased by too great a grounding in nuclear physics theory, or even electrochemistry, nor were they experts in any experimental fields that might lead them to be overly critical of their results. Storms and Talcott had begun their cold fusion work, as had most chemists, trying to duplicate the calorimetry, which they immediately found too difficult. They then ran 26 cells to study the loading of deuterium into the electrodes, and on June 1, started looking for tritium.

332 9 GARY

TAUBES

At the time news of their results hit the press, they had run over a dozen cells in various incarnations, one of which, labeled cell 30, seemed to

have been suddenly gifted with 80 times its initial endowment Ryszard Gajewski later called this the “‘major eruption which big commotion.” Cell 29 was up by a factor of three and a These numbers still made their results incompatible with

of tritium. created the half. either nu-

clear physics, Pons and Fleischmann’s excess heat, or Bocknis’s tritium,

or anyone else’s radiation measurements. But Storms and Talcott were not measuring for excess heat, nor is it clear whether they had any radiation detection equipment at all. Tritium, they would write, ‘“‘appears to be the least ambiguous and most easily measured product of the ‘Cold Fusion’ effect.” Although Storms and Talcott seemed to publish one draft paper after another, few of them made it to the world outside the laboratory. Their experimental protocol remained unknown for almost six months and, outside the cold fusion camp, for a year. The two researchers had taken a unique approach based on the proposition that the more deuterium they could induce to migrate into the palladium lattice, the more fusion they could generate. They planned to increase the deuterium loading by poisoning their palladium electrodes with sulfur, which is where their protocol got interesting. Storms and Talcott chose to make a sulfide using a technique that dates back to the nineteenth century and sounds like something that H. P. Lovecraft might have concocted. They melted the sulfur in the presence of household paraffin, then dipped the electrode into the resulting black goop. The result of the pre-treatment was a “‘black layer’’ covering the electrode. Nate Lewis later called this ‘‘black crud,” which was probably not the technical terminology. Unfortunately, this black layer/crud prohibited the deuterium from entering the palladium at all, which is not

surprising. Storms and Talcott were forced to reverse the current on the electrode, causing the crud to flake off the palladium and float quietly into the electrolyte. Whether it sank slowly to the bottom or floated to the top, they did not say. It was a complicated procedure. Then they re-reversed the electrolysis. Nine days later they ‘“‘decanted”’ the crud-filled electrolyte and apparently replaced it with fresh electrolyte. The next day they began measuring for tritium, and found an eighty-fold increase over what they expected. In their drafts, Storms and Talcott acknowledged that “‘this work is incomplete and leaves many questions unanswered.”’

One such question, of course, was What had

they expected? Although they tested the electrolyte before running, they apparently

BAD

SCIENCE

@ 333

did not test either the paraffin or the sulfur, so what could they expect? One week later they ran four new cells with the “Lovecraft” treatment, and none produced tritium; in September they ran four more. All eight cells produced low levels of tritium. Had they not expected to find tritium, they might have assumed that the paraffin, or even the sulfur, had been contaminated with tritium to begin with. Storms and Talcott did report that when they reversed the current on their cells, no additional tritium appeared. This argued for the conclusion that no tritium was made in the electrode, because if it had been, the current reversal would

have forced it out. But this was nit-picking. Storms and Talcott also wrote that even though these prickly questions remained, their results were supported by “such a large and consistent data base that reporting tritium production is warranted even before a full understanding of the process is available.” Eventually Storms and Talcott ran 150 cells and reported tritium production in 13 of them. They reported that 12 of these “produced” an increase in tritium from 1.5 to 3 times the original amount. Seven of these seemed to come from the same batch of palladium, which also argued strongly for contamination. The thirteenth was the Lovecraft cell 30. One of their very first cells seemed to be a control cell, with a nickel anode, and apparently they did not bother running another control until they had run over one hundred tests and garnered all their positive results. Even after two months of public and very vitriolic debate over the nonexistence of Pons’s controls, Storms and Talcott still didn’t believe it

was necessary to run sufficient controls. Like Pons and Fleischmann, they apparently believed that a cell that did not give the desired result—“‘‘did not work,” in the lingo of cold fusion—constituted a control.”’ In any case, Storms and Talcott ran six controls—only four with light water—

and 150 tests, which means their results crossed the line from dubious to

meaningless. Imagine the kind of pedagogical example that might show up ina high school aptitude test: a physician claims to have a cure for the common cold. He gives his cure to 150 cold sufferers and he gives a placebo to 5, which he says are his controls. The next day, 13 of those cold sufferers who took the cure are healthy. None of the 5 who took placebos is. What can be concluded? Nothing. How does the physician know that his 13 cures would not have returned to health without his expert help? He doesn’t. Hence the Food and Drug Administration would suggest, among other things, that the physician give the cure to 150 cold sufferers and the placebo to another 150. And that he try to keep the populations

334 HF GARY

TAUBES

of the two groups as identical as possible. If the former were all in Arizona and the latter in Boston, the test would be meaningless regardless of the results. If the former had all been sick for an average of a week and the latter for an average of two days, or four days, or maybe even six days, it would be meaningless. And so on. It’s a tricky business. What about those questions that the work left unanswered?

For instance, did Storms and Talcott know what would

happen if they ran 150 cells, say, with light water and palladium electrodes, or heavy water and gold electrodes, or even red wine and escar-

gots? They might also have observed slight, anomalous increases in the tritium content ofa dozen cells. In which case cold fusion, one supposes, could not be responsible for any results. Thus contamination would seem to be the most likely explanation. In fact, considering that the experiment was being conducted in a nuclear weapons lab, contamination seemed a much more reasonable explanation of the results than a heretofore undiscovered phenomenon that violates the laws of physics by a 100 quindecillion or so. Instead, Storms and Talcott chose to claim that “reporting tritium production is warranted even before a full understanding of the process is available.” The Los Alamos experiment, in Dick Garwin’s words, was really bad work. The New York Times ran the Los Alamos tritium story on June 27. This was, apparently, after Storms had his feet held over the coals. Under the headline, SIGNS OF ““COLD”’ FUSION ARE CITED, CAUTIOUSLY, the article by William Broad began, “‘Cautioning that the finding may prove incorrect, scientists at the Los Alamos National Laboratory in New Mexico said yesterday that they had found evidence of tritium.” It was not a bad way to handle the story. The Wall Street Journal didn’t run the story until nine months later, when the reporter, Jerry Bishop, obtained a copy of a Storms-Talcott preprint. The headline read COLD FUSION RESEARCH DISPELS SOME DousTs. Bishop wrote that because of the Los Alamos results, “doubts that earlier measurements of heat and tritium were artifacts, or products of scientific error, are fading.’” He did not quote Storms or Talcott, which suggests that the two were unwilling to have their soles toasted again. Bishop did quote John Bockris extensively, as well as Dave Worledge, a funding agent at the Electric Power Research Institute, which was underwriting Bockris’s tritium research. Ryszard Gajewski of the Department of Energy initially sided with Bishop and the Journal. Gajewski actually managed to give Storms and Talcott $330,000 from his Advanced Energy Projects coffers to continue their tritium work. He said DOE would be derelict in its duty if it didn’t

BAD

SCIENCE

#@ 335

fund them. This gift was then cited as confirmation of the tritium by The Deseret News, which quoted Bockris saying triumphantly, “It’s the end of the game—the beginning of a new tomorrow. The battle is finished.” It seemed as though an entire field of inquiry could be created with two researchers commenting on each other’s work helped along by a funding officer and a journalist who both happened to be carrying the same torch. The only item that would be missing would be the science.” As was true with many of the erroneous reports of cold fusion, it wasn’t the science of the Los Alamos tritium that was important, it was the timing.

10

SALT

LAKE

CITY.

ADMINISTRATORS

On June 28, one week after the Los Alamos news and one day before the state’s fusion panel assembled to decide on the fate of the $5 million, the University of Utah announced that General Electric had finally agreed to be a partner in its cold fusion future. It was the promise of this agreement—not to mention the heat from Bob Huggins, bursts from Milt Wadsworth, and tritium from John Bockris and Ed Storms and Carol Talcott—that gave the U faith to push for confirmation. Nonetheless, the agreement very definitely favored GE, which had not been the case back in April. Originally, GE had been courting Utah. Now it was the other way around. With the partnership a fait accompli, all GE had agreed to do was involve its electrochemists in the cold fusion project, and provide money, if certain bench marks could be achieved. Specifically, GE agreed to put up $500,000 if the U could prove a cell was generating a tenth of a watt of excess power; $5 million for ten watts of excess power.”? Since May, GE’s electrochemists had been allowed to examine Pons’s calorimetry in his laboratory. Pons had even let them dismantle a working cell and take it back to GE headquarters in Schenectady, New York. General Electric also sent Steve Spacil, an expert in thermodynamics, to Utah for three weeks to wait for heat bursts, but Pons had left for Britain

to visit Fleischmann, and he instructed Mark Anderson not to let Spacil into his lab in his absence.

336 @ GARY

TAUBES

Pons returned a few days before Spacil’s scheduled departure. Just after Spacil left, Pons reported that one of his cells had indeed been generating huge amounts of heat, but nobody had realized it. (Before Pons returned to Utah, he told Clive Cookson of the Financial Times that his Salt Lake

City colleagues had been faithfully faxing him the latest readings from his experiments, so maybe Pons had known there had been a heat burst but preferred not to let on. Or maybe there had been no heat burst at all.)°° However, Wadsworth’s cells did emit a healthy burst of heat while Spacil was still around. Thank God he was there, Wadsworth said; it was like

seeing a UFO and having a witness to validate one’s sanity. Later Hugo Rossi wrote in a progress report on cold fusion that the “GE thermodynamicist concluded that those data [from Wadsworth’s lab] cannot be explained by any error in measurement, nor any chemical phenomenon.”’ But GE never did provide money for cold fusion research, so there seems to have been a miscommunication. As the DOE experts had done, Spacil suggested that both Pons and Wadsworth install redundancy in their electronics. A month later, when the U inaugurated its National Cold Fusion Institute, Chase Peterson claimed that GE was in for a “‘sizable’’ invest-

ment, which would increase, not surprisingly, if the fusion experiments were successful. However, he gave no details. Jim Brophy, indomitable as ever, said that they had been sworn to secrecy on the agreement because GE officials had been ‘‘offended’”” when The Salt Lake Tribune published the details of Hugo Rossi’s report on the company’s fusion research. That report was very illuminating, although it was hard to see why GE had taken offense. It summarized recent cold fusion results, including those of GE’s analysis on the working fusion cell that had been taken to Schenectady. This section was convoluted, to say the least: Scientists at General Electric, after long and careful study, some of it here with Pons in his laboratory, conclude that the basic calorimetric theory of Pons and Fleischmann is correct and shows excess energy. (They have some concerns about the accuracy of the calibration, and have made a

correction in their own preparation of the data. With this correction (amounting to 15%) F&P data show significant excess energy.) They have reproduced the experiment in Schenectady and have obtained excess energy at about the 15% level, thus indistinguishable from 0 with their

correction.

What Rossi was trying to say, and nearly succeeding, was that Pons and Fleischmann had reported 15 percent excess heat. The GE chemists had

BAD

SCIENCE

@ 337

run the cell themselves and found that if Pons and Fleischmann’s calorimetric assumptions were used, the cell did indeed produce 15 percent excess heat. However, GE believed that there were various errors in the Utah calorimetry and that they accounted for the 15 percent “excess heat,” thus making it “‘indistinguishable from 0.”’ So an independent party had finally been allowed to examine an ‘active’ cell from Pons’s lab and found it to be in perfect energy balance. No excess heat.*! Once Pons discovered that the GE scientists were critical of his calorimetry, he refused to discuss the experiment with them any longer. Ian Cumming then suggested that as long as GE was only interested in providing negative input, perhaps their role should be limited to providing ‘‘ongoing quality control.” Meanwhile, Utah flew in Huggins and Bockris for the cold fusion confirmation hearing on June 29. Huggins told the committee that he had confirmed the production of excess heat from electrolytic cells, but he wasn’t ready to swear it was fusion. Bockris said the same for the tritium, heat, and neutrons. The two must have appeared downright cautious, and the advisory panel, composed mostly of nonscientists, was eminently capable of being snowed by two such seemingly prestigious researchers. Bockris suggested that the committee consider distributing some of the $5 million to researchers outside the state of Utah. He remarked to Rossi that there were limits to what Utah could expect from Fleischmann and Pons, and the best thing to do with that money was send it all to Texas. Nonetheless, the advisory panel decided to wait yet another two weeks before voting on whether cold fusion had been confirmed. In the meantime, Pons appeared in living color in The Deseret News standing next to a mysterious device about the size of a bread box. It was captured for posterity on the network by a civic-minded Utah computer scientist: The cylinder is somewhere between 1 and 1% feet high, with a thick, shiny metal base and a red screen, capped with metal and something else (plastic?). Something dark in the center can be seen dimly through the screen. The top has a couple valves plugged into it and two clear plastic hoses are attached to the valves; black and red wires run into the top from an unseen source (the usual car battery?). A digital voltmeter is monitoring the two leads. Pons’s hand appears to rest casually upon a complicated glass gizmo which looks like it may be part of the innards of the red cylinders. It was a “mini-boiler,”’ according to Pons, and it generated twenty

times more energy than it consumed. It was also an attempt to develop

338 9 GARY

TAUBES

a cold fusion technology parallel to the basic science, as prescribed by Ira Magaziner. It could also be viewed as a public relations stunt to sway the committee into voting for confirmation. Pons told the reporter JoAnn Jacobsen-Wells that the device “‘wouldn’t take care of the family’s electrical needs, but it certainly could provide them with hot water yearround.” Jacobsen-Wells then explained for her slower readers: “Simply put, in its current state it could provide boiling water for a cup of tea.” And Jim Brophy, who had a way of distilling the absurdity out of a situation, said, “We

can boil water, but I’m not sure it’s for making

coffee or taking a shower.” This prompted Steve Jones to warn that Pons and Fleischmann could not have it both ways. If cold fusion produced tritium in the quantities that Bockris, for one, had reported, then the miniboiler would generate

dangerous amounts of tritium if it worked, which Jones didn’t believe for an instant. Jones told reporters that he personally was a victim of tritium poisoning; he had once inhaled the noxious stuff in a minor research accident. He was lucky that tritium stays in the body for a very short time, but he wondered whether his exposure could have contributed to the fact that three of his children born after this accident had minor birth defects. ‘‘It’s a terrible hazard,”’ he said, “‘it really is.”’

11

PALO

ALTO.

WASHINGTON,

D.C.

A hypothesis or a theory is clear, decisive and positive, but it is believed by no one but the man who created it. Experimental

findings,

on

the other

hand,

are

messy,

inexact

things, which are believed by everyone except the man who did the work. HARLOW SHAPLEY, former Harvard astronomer

On July 6 the Department of Energy’s panel of experts visited Bob Huggins’s lab at Stanford, where he informed them that he had no cells producing excess heat. Neither, for that fact, had John Bockris nor John Appleby nor Stan Pons nor Milton Wadsworth when the panel came around. These cold fusion researchers were beginning to remind the DOE experts of Uri Geller, who couldn’t do his mentalism while anyone

was watching closely.

BAD

SCIENCE

§ 339

Huggins said he had no cells running because their old building was being renovated, and they had just that week moved to a new one. This was true, but it still provoked cynical remarks from the panel members, to whom it sounded like another excuse. That set the tone for the meeting. Huggins had already been warned by Pons and Bockris that the panel would be aggressive and the experience unpleasant. The experts could not understand why Huggins was so uncritical of the results of his own experiments, and Huggins and his researchers couldn’t understand why the DOE people were so negative. ‘““They just wouldn’t believe anything you told them,” Huggins said. The panel found Huggins’s experimental setup the least impressive of any they had yet seen. He had one little room that Will Happer described as pathetic, with bits and pieces of apparatus, some of it not even wired together properly. That Huggins’s experiment was riddled with artifacts seemed less bizarre, however, than that he didn’t appear to care. In meetings he displayed data without any indication that measurement errors existed, and in his lab techniques, he seemed unconcerned about

the possibility of such errors. Happer recalled that the panel members noticed a suspicious correlation in Huggins’s data between excess heat events and weekends. “On weekends,” Happer said, “industries close down, and it’s common for live voltage to rise a little bit. Huggins didn’t know about that. It’ll go from 115 to 120 volts, which is a big effect, comparable to the effects he was looking at.” Larry Faulkner recalled that the Stanford experiment was susceptible to AC power fluctuations. The researchers measured the direct current power going into the cell but did not take into account the alternating component of the voltage. These could result in errors of the size of their excess heat effects, but again they seemed unaware that the fluctuations existed. Faulkner also remarked that the Stanford electrodes had bare surfaces, which meant the evolving gases could recombine on the bare platinum and palladium, releasing large amounts of heat back into the cell. Even Pons and Fleischmann had shielded their metal with glass tubing and tried to estimate how much gas was escaping and thus how much was recombining. Huggins’s researchers professed not to understand why this would make a difference. ‘“They didn’t even address that issue,” Faulkner said.

Al Bard noted that the thermocouples in the Stanford cells had been inserted directly between the two electrodes, so the resulting electric field would affect the temperature readings. Huggins’s researchers disagreed, so Bard had them take one cell, switch the power on and off, and

340 8 GARY

TAUBES

watch the temperature reading, which changed by a tenth of a degree. It was a small effect, but still one to be avoided. Huggins hadn’t noticed the effect until Bard brought it to his attention. None of this constituted proof that Huggins had not electrochemically induced nuclear fusion—or, as he would put it, that he hadn’t confirmed

the excess heat production—only that he had no evidencé to make that case. Happer later remarked that Huggins came into cold fusion with a good reputation, unlike Bockris, for example. The moment certain people heard that Bockris was involved, said Happer, they discounted cold fusion: “Huggins elicited another response, maybe there was something to it. Having met Huggins, of course, my reaction to him from now on will be the same as [to] Bockris.”’ When Huggins was later asked how he could express such conviction in his data while admitting that his only personal role had been to fetch Peking duck for his researchers, he responded only that they had begun the experiment on April Fools’ Day. This may have been his peculiar sense of humor. Steve Crouch-Baker, who did the Stanford calorimetry, suggested that Huggins’s faith sprang from his desire to procure funding. ‘““You have to be very positive about what you do,” he said. “If you’re the Department of Energy, and I say, “Well, I might have something interesting and I just need to confirm it, I’m not sure,’ you're not going to give me any money. Obviously if I come in and say, ‘Look, this is the way it is. We saw this, this, etc.,’ I am far more

likely to get some money. That’s just human nature.”’ It was also, once again, Pascal’s wager. Al Bard came away from the Department of Energy laboratory visits deeply depressed. The low quality of the science in these labs had been a revelation. Probably the most critical shortcoming, as he saw it, was that the proponents of cold fusion had no understanding of the importance of statistics. They would run asingle cell, obtain a positive result, and claim confirmation of cold fusion. It was as if, when confronted with

salvation they could ignore common sense, as if nature wouldn’t possibly lie to them on such an important matter. They didn’t realize that they had to account for measurement errors and statistical fluctuations. To do that, they had to understand the minute details of their experiments, and repeat the experiments again and again. In any given lab, explained Bard, the results of a series of calorimetry experiments would fall on a bellshaped curve, some with excess heat, some with a heat deficit, all dis-

tributed around zero:

BAD

Considering the measurement errors inherent in calorimetry, a small number of all the experiments should have been expected to show excess heat of, say, 6 percent or more. If these—happening only “‘sporadically”’ and thus “‘not reproducible,” as Bockris would put it—were claimed to be evidence of fusion, and all the rest of the negative experiments were ignored, the result would be the phenomenon of cold fusion. Curiously enough, what Bard learned to appreciate from cold fusion was that the same statistical lesson could be applied to real life. Cold fusion could be understood in the following way: on March 23, thousands of scientists began simultaneously doing identical experiments. These scientists ranged from the very best around to the very worst, which means their level of talent could also be plotted on a bell-shaped

curve:

4

e.*

‘+

2 : ;

> Ate

we

oe

Die

*

hus vr

| leased tung i

Mat

7

,

;

£

‘5 ee“is aes wit! -

‘i

9")

~—~S

,

:

ates pang won

ata2

%

wal

©

meri?

? devi

vo

hae

a

a

«

4

7

eS

WG. ue

2)

>

z

:

>. ae sel

ie

on

'

eS

«

a

So"

y lines ay > Shi li 4 ¢ Seat eb —— we fhe a AA SRE es et. is

ee

a

er G pee ete = Se. ve. meet AAT

Ane glee eee

Pars

7 Jing

artis

;

«

;

6

on:

x

: 2

ei

ves

> j

9

} ab

-



; pest ;

ten rag

es.

iy

reruns

— -

SRE.

q

-.

es Y HENS “ on tue.