Category Something New Under the Sun

Something New. Under the Sun

When the story of our age comes to be told, we will be remem­bered as the first of all men to set their sign among the stars.

—Arthur C Clarke, The Making of a Moon, 1957

A new age was dawning, in which the organized brain power for military and civilian science and technology was the dearest national asset.

—Walter McDougall,… the Heavens and the Earth: a political history of the space age, 1985.

W

hen the Alfred P. Sloan Foundation decided to sponsor a series of histories of technology in the 1990s, they asked for proposals about technologies that have had a significant impact on the twentieth century. For some years, I had written news and features about space and had been hooked by the glamour of space exploration. Which aspect of the field would, I wondered, best fit into the Sloan’s proposed series?

It seemed to me that the answer was navigation, weather, and com­munications satellites—that is, so-called application satellites. The National Academy of Engineering has said that of all the technological achieve­ments of the second half of the twentieth century, these satellites are second only to the Apollo moon landing. Application satellites have a stealthy, silent influence on our lives. Most of us would notice them only in their absence. But then we would notice. There would be no early warning of hurricanes, no satellite data for the computer models that predict weather. There would be no instantaneous communication to and from any part of the globe, no satellite TV, and no navigation in bad weather. It would be a more dangerous and expensive world.

So I submitted a proposal. I wrote blithely of a history of every kind of civilian application satellite, from every country, from before the launch of Sputnik up to the 1990s. My book was also to encompass the critical supporting technologies of launch vehicles, electronics, and computers. Somehow, I won the grant.

After a few months of research, I discovered that I had known little about the subject and that it was full of apocryphal tales from imperfect memories. It took about three years to track down participants and locate archives, company records, and small pockets of papers kept by people when they retired. I had to find out what was classified and what wasn’t, what people thought was classified even when it wasn’t, and what, in gen­eral terms, might be in the genuinely classified material.

Not surprisingly, my original proposal was of absolutely no use. Its main fault was that it was about civilian application satellites. But navigation satellites were developed for a purely military purpose. The early history of weather satellites is inextricably intertwined with that of reconnaissance. And the decision that led to Syncom, the precursor of Early Bird, the world’s first commercial communications satellite, owed much to the mili­tary’s urgent need for improved global communication.

So the word civilian was the first thing to be excised from my con­ception of the book. It was followed by a ruthless culling of the 1990s, the 1980s, the 1970s, most of the 1960s, and satellites developed outside the United States. Finally, all but a few of the early American satellites fell by the wayside. Launch vehicles, electronics, and computers survived by the skin of their teeth, and only insofar as they demonstrated the limitations and difficulties surrounding those designing the early satellites.

What is left gives a flavor of yesterday’s technology, which is our own technology in embryo, and a technology that has shaped our world. The book excludes many people, which is a shame but inevitable if it is to be readable.

The title, Something New Under the Sun, is a play on the biblical say­ing that there is no new thing under the sun. It was coined by Bob Dellar, an amateur astronomer who led a group of “Moonwatchers” in Virginia in 1956. The task of the Moonwatchers, who were scattered all over the world, was to track the satellites that the United States and the Soviet Union were planning to launch during the International Geophysical Year of 1957/58. Mr. Dellar is now dead, but Roger Harvey, who was sixteen at the time, was one of Mr. Dellar’s group, and he mentioned the phrase in a parking lot in northern Virginia while we inspected the telescope he had used to search for Sputnik. I asked if I could purloin the phrase as the title of my book, and he said yes.

It is an apt title, because satellites were, literally and figuratively, something new under the sun. The pioneers who designed the first satel­lites admit cheerfully that they hadn’t a clue what they were doing or what they were up against. Their launch vehicles blew up, their electronics were unreliable, guidance and control were primitive, the world was just turning from vacuum tubes to transistors, and those transistors didn’t always work. The list of things they didn’t know and that failed goes on and on and those things are, of course, the reasons why those early participants in the space age were pioneers.

The only non-American participant who is discussed at any length in the book is Sergei Korolev, the mastermind of the Soviet Union’s space program who was responsible for the launch of Sputnik. He was an extra­ordinary man of extraordinary tenacity, who at great personal cost survived Stalin’s paranoid and casual cruelties. Despite his contribution to the Soviet Union’s Cold War armory, some tribute seemed called for, and the so-called chief designer of cosmic-rocket systems has the introductory chapter to himself.

Sputnik, according to the historian Walter McDougall, sparked the biggest furor in the United States since Pearl Harbor. The satellite was Korolev’s baby, and it was launched as part of the International Geophysi­cal Year.

The IGY was the brainchild of Lloyd Berkner, a leading American scientist. While scientists were still in the early stages of planning the IGY, President Eisenhower announced that the U. S. would launch scientific satellites as part of its contribution to the IGY. Within days the Soviet Union made a similar announcement.

It seemed at the time that the White House had bowed to pressure from industry, scientists, and the military. But recent scholarship suggests that President Eisenhower hijacked the IGY and made it, in the words of U. S. Air Force historian R. Carghill Hall, the stalking horse for the admin­istration’s plans for reconnaissance satellites.

As is so often the case, there were many people to whom it didn’t matter at all why what happened, happened. The space age had opened, and the pioneers of navigation, weather, and communications satellites were ready.

At the Applied Physics Laboratory (APL) of the Johns Hopkins Uni­versity, in Maryland, Bill Guier and George Weiffenbach listened to Sput­nik’s signal. Within the week, they were developing an approach to orbital determination that broke with a centuries-old tradition. Before Guier and Weiffenbach’s work, the technique was to measure the angles to heavenly bodies and to determine orbits from those values. The scientists of the IGY had an elaborate optical and radio observational system in place for mea­suring the angles to satellites. Guier and Weiffenbach measured changes in

frequency and developed computational and statistical techniques that at the time seemed to be coming from “left field.”

Their boss, Frank McClure, adapted their techniques to form the basis for the Transit navigation satellites. In turn, Transit severed links with millennia of esoteric navigational rituals by providing mariners with com­puter readouts of latitude and longitude. Transit was developed because the Special Projects Office of the Navy needed some way to locate its Polaris nuclear submarines with greater accuracy than possible with existing methods. But the system went on to serve surface fleets, merchant vessels, the oil industry, fishing fleets, and international mapping agencies.

In the Midwest, Verner Suomi, of the University ofWisconsin, heard a lecture about the IGY and proposed flying an experiment to measure the radiation balance of the earth, a value of fundamental importance to meteorologists. The experiment set him on the path to earning the hon­orary title of father of weather satellites. These satellites took twenty years to find widespread acceptance among meteorologists. Because they rely on similar technology to that of reconnaissance satellites, they have a murkier history than that of satellite navigation.

In New Jersey, John Pierce (known in science fiction circles at the time as J. J. Coupling), of Bell Telephone Laboratories, played the pivotal role in the early days of the development of commercial communications satellites. He was swiftly challenged by Harold Rosen, of the Hughes Air­craft Company On the title page of his book How the World Was One, Arthur C. Clarke calls Pierce and Rosen the “fathers of communication satellites.”

Guier, Weiffenbach, McClure, Suomi, Pierce, and Rosen—these were the Edisons and Marconis of satellites for navigation, meteorology, and communication. Pierce and Rosen were rivals; Weiffenbach heard Pierce lecture and learned some things about satellite design; Suomi sought Rosen’s help when he was trying to persuade NASA to fly another of his experiments; Weiffenbach met Suomi in India. They were not all close friends, but the American space community of the late 1950s was small and intimately connected. The same mysteries faced them all: the unimag­ined complexity of the earth’s gravitational field, the unknown space envi­ronment and the radiation belts. All but Rosen benefited directly from the IGY. All were involved in projects that ultimately became the work of hundreds.

Something New Under the Sun is about the ideas of these men and the global, national, and local influences that shaped them.

In the case of Transit, most of the primary source material comes from APL and some of that material could be extracted for me only by people with the necessary security clearances, so there may well be things I am missing that I do not know about. The story is told through the eyes of APL, even though I have tried to set it in context. I’m sure someone view­ing the story from outside APL would have a different tale to tell, but as written, I hope it gives a sense of what it was like to develop a satellite sys­tem in the late 1950s and early 1960s.

Meteorology satellites were more difficult to write about because the story is intricately linked with the change from art to science that meteo­rology was undergoing in the 1950s and because much of the primary source material was still classified when I was writing. But key participants helped to steer me through a sea of partial information. Verner Suomi is one of several who played a critical role in the early days, and perhaps more should be said about the others. If anyone writes the story at greater length, perhaps more will be said.

The early days of communications satellites are described mainly from the point of view of the Bell Telephone Laboratories and the Hughes Aircraft Company. Much of the text is based on material I collected from the archives and company records of AT&T and Hughes, supplemented by interviews and by other documents that participants passed on to me.

Echo and Telstar are names that still bring a flash of recognition to some faces. They are part of this book because John Pierce, whose ideas were important in many ways, was involved either directly or obliquely with them and because they highlighted AT&T’s plans for global satellite communication, which raised antitrust concerns that shaped American policy in this strategically important new field. For these reasons, I have concentrated on Echo and Telstar rather than the NASA-sponsored Project Relay.

The most famous satellites that were based on Rosen’s initial ideas were the Syncom satellites and Early Bird. These opened the era of com­mercial communications satellites.

Navigation, meteorology, and communication—ancients arts that have become sophisticated science and technology. The application satel­lites that ring the earth did much to help in that transition. Our social institutions and expectations are changing rapidly. Via satellite, we can remotely diagnose illness, watch from our living rooms while “smart” weapons shatter their targets, or track the development of major hurri­canes for weeks before they threaten our coasts. The impact of satellites on our lives has scarcely begun.

Storm Patrol

People have told me I’m a wonderful salesman, but it took all of my salesmanship while I was in Washington [to persuade NASA to fly the spin-scan camera].

—Verner Suomi to author, May 27, 1992

The spin-scan camera was a giant step. It gave you a view you didn’t have before.

—Robert White, former president of the National Academy of Engineering, to author, 1992.

I

n his long, narrow office at the University of Wisconsin, with awards hung on the walls (others are stuffed in drawers in the basement), Verner Suomi recalled the first spin-scan camera that he and Bob Parent proposed to NASA in the fall of 1964. “It was disgustingly simple. The stuff on the ground that you need to put the pictures together, that was not so simple.” The camera’s job was to continually monitor the weather over one portion of the earth’s surface.

The space-based elements of the idea were, indeed, conceptually straightforward: a spinning spacecraft in geostationary orbit,[12] a telescope, a camera, and a data link with the earth. The practicality was a little more difficult. “But,” said Suomi, “one of the advantages was that we didn’t know what the problems were, so they didn’t hold us up.”

Suomi and Parent’s first proposal for a spin-scan camera, dated Sep­tember 28,1964, was a hastily thrown together three and a half pages of text and two pages of very simple diagrams. Parent was the electronics expert. Their proposal was called “Initial Technical Proposal for a ‘Storm Patrol’

Meteorological Experiment on an ATS Spacecraft.” Like other meteorolo­gists at the time, Suomi wanted to take advantage of the 22,300-mile-high geostationary orbit, in which a satellite stays in the same position, more or less, with respect to the earth and thus “sees” the weather moving under­neath. Polar-orbiting satellites, by contrast, see successive snapshots of the weather in different places as they move through their orbit.

A geostationary orbit, however, is a long way away from the earth, so Suomi and Parent described a telescopic camera that would enlarge the distant image. Since only a small part of the earth would fall within the field of view, some method was needed of scanning in the east-west and north-south directions to build up an image of the earth’s surface. The satellite on which they hoped to mount the spin-scan camera would be spinning at a steady 100 rpm, and thus automatically would scan a line from east to west. After each revolution, the camera would shift its field of view slightly to build the full picture of the earth’s disc. Over the years, several electronic and mechanical methods of achieving movement in the north-south direction were explored.

Suomi and Parent thought that the image could be built over ten minutes from one thousand scan lines, giving a resolution at the subsatel­lite point of six nautical miles. In their second, ten-page proposal to NASA a year later, the camera, which was designed cooperatively with the Santa Barbara Research Facility of the Hughes Aircraft Company, had an image built from two thousand lines and thus an improved spatial resolution.

Today’s technological descendants of the first spin-scan camera scan sixteen thousand lines in thirty minutes. During severe storms they can build more frequent pictures of smaller regions. They observe in the infrared. Each radiometric reading is assigned a color, and a false-color image is created. From these images, meteorologists, infer wind speeds, which are particularly important for modeling atmospheric conditions in the tropics (within thirty degrees of latitude north and south of the equa­tor), where the temperature differences are too small for satellite sounders to make distinctions.

Despite the greater spatial resolution of today’s satellites, Suomi, talk­ing in 1992, was not happy about the thirty-minute time interval between photographs. In his opinion, the ideal interval is the ten minutes that he and Parent first proposed in 1964 because in that time very little change in the weather and very little detail of an evolving weather pattern is lost.

In that first proposal, Suomi wrote, “The object of the experiment is to continuously monitor the weather motions over a large fraction of the earths surface.” He and Parent envisaged a camera that would observe the earth between fifty degrees of latitude north and south, which would, of course, encompass the meteorologically all-important region of the tropics.

Suomi quoted results from his radiation balance experiments on Explorer VII and several of the TIROS satellites to make his case, writing that the amount of radiation reflected from the tropics was lower than expected, even though the total outgoing radiation from the earth was close to earlier estimates. Thus, more heat than previously thought was being transferred from tropical to polar regions.

The questions meteorologists needed to answer were, How was that heat transfer achieved, and how did it affect global circulation of the atmosphere? They had few observations with which to work because the tropics—the “boiler,” as Suomi wrote, of the giant atmospheric heat engine—which cover about half of the earth’s surface, are eighty percent ocean. The polar orbiting TIROS satellites did not help much. Those satellites spent only about fifteen minutes traversing the region as they headed north (similarly for the southward journey) in their orbit. There was a gap of twelve hours before the spacecraft was next above the same subsatellite point. In the tropics, where weather patterns develop and dissi­pate in far less than twelve hours, the result was that the TIROS satellites did not provide observations of the complete life cycle of a typical tropical storm. Instead, meteorologists inferred the progress of a “model” storm from observations of different storms in different places at different stages of their development.

Yet these storms, including hurricanes, are one of the mechanisms by which the “boiler” of the “atmospheric heat engine” redistributes heat around the earth. The rationale of the spin-scan camera was to provide data that would allow meteorologists to explore these mechanisms.

It took more than a decade for meteorologists to find effective ways of exploiting the spin-scan camera, but eventually inferences of wind speeds in the tropics improved atmospheric models.

Bob Ohckers, an electronics technician who joined Suomi’s group in 1967 from RCA, said that Suomi initially wanted to measure the winds from the displacement of clouds between successive images. “We’d get one image (an 8 by 8 transparency) in a frame and superimpose a second image taken twenty minutes later. First, we’d line up the geographical points in the two transparencies, completely ignoring the clouds. Next we’d shake the images in a frame until the clouds from the two images were superim­posed on one another and the geographical features were displaced. You could tell when the clouds coincided because the light shining through from below was at it dimmest in those places. Then you would measure the x and у displacement of the clouds.” The method worked, but it was impractical, and the department’s software group came up with a better way of doing the same thing. When Suomi saw the results of the software, he dropped the mechanical approach without a backward glance.[13]

Although meteorologists in the early 1960s were keen to observe the earth from geostationary orbit and plans existed on paper for a geostation­ary meteorology satellite, there was a problem. “No one had any idea,” recalled Suomi, “about how to get the blooming thing up there.”

Then Harold Rosen, Donald Williams, and Tom Hudspeth, of the Hughes Aircraft Company, came up with the engineering concepts that made attaining geostationary orbit both economically and technically feasible at an earlier date than anyone had thought possible. It was an advance that was to be a key factor in opening up the multi-billion-dollar business of civilian communication satellites in the mid 1960s, but a description of a NASA satellite based on the Hughes design also fired Suomi’s imagination. It was called, prosaically enough, the Application Technology Satellite-I. ATS-I was to carry an experimental communica­tions payload with sufficient bandwidth to transmit a TV channel.

Suomi’s attention was caught by the simplified block diagram that accompanied the article describing ATS-i. It looked to him as though the satellite should be able to carry a small camera and that there would be sufficient bandwidth to carry its images back to Earth.

During July and August 1964, Suomi elaborated his ideas, and he and Parent hastily put them into their September proposal to NASA.2

Storm Patrol

Earlier in the year, Suomi had completed a brief stint as chief scientist of the Weather Bureau, working for Robert White (who later became presi­dent of the National Academy of Engineering, retiring in 1995). “Wouldn’t it be nice,” Suomi now asked White, “to beat the Russians into space with a camera viewing the weather from a geostationary satellite?” Seven years into the space age, many space scientists and engineers still felt they needed to regain the technological initiative from the Soviets. White’s practical response was to grease the bureaucratic wheels for Suomi, who, as with the International Geophysical Year, was making a belated entry into a satellite program.

NASA at first told Suomi that the spacecraft would not be stable enough for his camera. Suomi called Rosen at Hughes, who, incensed by the comment, made his own phone calls to NASA.

In the meantime, Suomi presented his and Parent’s ideas to govern­ment officials and industry representatives, including TRW and the Santa Barbara Research Center of the Hughes Aircraft Company. Both compa­nies invested their own resources to investigate the concept. Several data processing issues had to be solved. For example, the camera was being designed to have a precise geometry, and the geometry of the resulting image had to be preserved after processing. Second, from geostationary orbit, the Earth occupied only about 16 degrees of the camera’s 360 degree field of view (because it was rotating). So the camera would be recording images of the earth for only about a twentieth of each revolu­tion, and the signal would take up twenty times more bandwidth than was needed to relay the image data. There were questions, too, about the impact of camera distortion and about nutations of the spin axis (preces­sion).

NASA backed the proposal in time for the camera to fly on the ATS-1 spacecraft. Suomi kept the technical authority for the project at the University ofWisconsin but subcontracted the physical construction and final engineering to the Santa Barbara Research Facility.

Some years later, Hughes filed a patent on the spin-scan camera, but Suomi opposed them, supporting NASA’s claim to the patent because it was the agency that had funded his work and because Suomi believed that the validity of the Hughes patent claim rested on his ideas. NASA, which would be less fortunate during a later patent dispute with Hughes about crucial elements of the Williams, Rosen, Hudspeth satellite design, won the dispute. Nevertheless, as Suomi said some years later, Hughes engineers made important contributions to the development of the camera, and, he added, . Hughes built the camera, so in a manner of speaking, they reduced the idea to practice.”

Suomi almost missed the launch of ATS-1. He had forgotten to do the paperwork for his security clearance, but a colleague interceded for him. Suomi said his most exciting professional moment came when the first image of Earth’s disk ever taken from space appeared on an oscillo­scope. The aim of the spin-scan camera had been to have weather imagery available to meteorologists in real time. That did not happen immediately. The first printed images from the spin-scan camera on ATS – і were ready four or five days after the launch. Suomi was scheduled to give a lecture at the American Meteorological Society. He said, “I had a whole bunch of negatives, and I tried to line these up with one another. I put a pin through, and I made a “movie.” I gave my talk and ran the movie. They thought it was wonderful to see the clouds moving.”

They had, in fact, seen the first ever animated picture of the earth’s weather—the primitive precursor to the pictures that appear today on television weather forecasts. There was still a long way to go before the technology would be regarded as mature, but one of the two most signifi­cant classes of instrument (the other was the sounder) that would facilitate that process was aloft. And it was mounted on a satellite that was the tech­nological kin of Early Bird, the world’s first commercial communication satellite

Chapter nine: Kersher’s Roulette

Comments about Richard Kershner (page 91), his approach to the job of team leader and to engineering, are based on the views of different Tran­sit team members.

The First Transit Proposal, 4 April 1958, (APL Archives) gives details of the satellite and incorrectly suggests that the ionosphere might be the biggest problem facing satellite navigation (pages 92-94).

Limits on orbital configuration and its relationship to ground stations (page 94) are from interviews with Guier and Weiffenbach.

The section in this chapter on the search for longitude had a number of secondary sources:

John HarrisomThe Man who Found Longitude, by Humphry Quill (Baker, 1966).

History of the Invention by John Harrison of the Marine Chronometer, by Samuel Smiles (Press Print).

Memoirs of a trait in the character of George III of these United Kingdoms, by John Harrison (W Edwards, 1835).

John Harrison and The problem of Longitude, by Heather and Mervyn Hobden (Cosmic Elk, 1989).

“The Longitude,’’ an essay by Lloyd A. Brown in volume two of The World of Mathematics, edited by James R. Newman (Tempus, 1956).

Kershner’s trips to the Pentagon (page 97) were remembered by both Guier and Weiffenbach. Though there is no written record of these trips at APL, he presumably had to go back and forth several times.

Transit on Discovery is mentioned several times in memos, letters, and progress reports of Transit in the APL archives (page 98), and various members of the team explained that it was part of DoD efforts to deter­mine Earth’s gravitational field and thus, of course, the forces that would act on a ballistic missile in flight.

The details in pages 98 to 104 were extracted from numerous reports and memos in the APL archives, from interviews with the Transit team mem­bers, and from memos and papers that Henry Elliott had kept.

Prologue

PrologueThe

T

he scientific mind is a curious thing. It probes what others take for granted, including, on one night in 1950, a multilayered chocolate cake. Some of America’s brightest scientific minds were focused on that cake at the home of James Van Allen, who was to become famous as the discoverer of the earth’s radiation belts and who was hosting a dinner for the eminent British geophysicist Sydney Chapman. With admirable atten­tion to detail, the collective scientific intellect verified that the cake had twenty-one layers. That cake did much to put the scientists in the kind of mood from which expansive conversation flows and big ideas are born.

Led by Lloyd Berkner, the talk turned to science. Berkner was a charismatic individual, head of the Brookhaven National Laboratory and a veteran of one of Admiral Byrd’s Antarctic expeditions. Like his fellow diners, Berkner was fascinated by the new insights into the earth’s environ­ment that instruments aboard V2 rockets captured from Germany had been providing since 1946. These high-altitude sounding rockets were invigorating the earth sciences. Was the time right, asked Berkner, to pro­mote an international effort in geophysics, one that would exploit new and established technologies in a world-wide scientific exploration to illu­minate global physical phenomena?

This conversation was to lead to the establishment of the Interna­tional Geophysical Year of 1957/58. The IGY was the enterprise that by inadvertently dovetailing with a key element of President Eisenhower’s national security policy—establishing the freedom of space for reconnais­sance satellites—was to become the cradle of the space age.[1]

At Van Allen’s dinner party in 1950, Berkner and his fellow diners were not considering spacecraft, though they all knew of the imminent technical feasibility of launching satellites and were to play important roles in the opening years of the space age.

Berkner’s suggestion for a geophysical year captured Chapman’s attention. These two agreed to “talk the idea up” among their many con­tacts in the international scientific community. Chapman sent an account of the dinner party to the journal Nature. Within a few years, Berkner and Chapman had secured enough interest to win the backing of the Interna­tional Council of Scientific Unions for the IGY. Chapman became presi­dent of the IGY and Berkner the vice president.

Scientists had already established the idea of international collabora­tion in 1882 and 1932 when they had undertaken to study geophysics from the earth’s poles. Expanding the concept of the International Polar Years to encompass the whole earth was, agreed the diners, a good idea, and the best time for such an effort would be between July 1, 1957, and the end of 1958. They chose the dates to coincide with a period of maxi­mum solar activity, when there would be much to study.

During solar maxima, which occur once every eleven years, the sun throws out huge flares of matter and energy more frequently than at other times, adding peaks of intensity to the solar wind that is always racing through the solar system. Numerous terrestrial effects result. The northern lights, for example, move to lower latitudes than usual as more charged particles precipitate along magnetic field lines into the ionosphere, an area of charged particles at altitudes of between 60 and 1000 kilometers above the earth’s surface. Studying the ionosphere was to account for a significant portion of the IGY, including science that was important for long-range communication, missile development, and, as it turned out, the space age.

Others in 1950 wanted to set up observation posts in places where meteorological data were sparse. Some wanted to organize expeditions to places where they could observe total eclipses. And some, like Lloyd

available to McDougall, that the Eisenhower administration wished to finesse a satellite into orbit in order to establish the freedom of space. See also The Eisenhower Administration and the Cold War; Framing American Astronautics to Serve National Security; by R. Cargill Hall (in Prologue, Quarterly of the National Archives, spring 1996). Cargill Hall’s article, which is based on additional, recently declassified documents, makes the argument more explicitly that the IGY and national security policy were linked.

Berkner, were intrigued by the physics and chemistry of the upper atmo­sphere.

So, with endorsements from the international scientific community, the participating countries got down to business. Each needed a detailed research plan. And they needed money.

In the U. S., scientists turned for leadership to Joseph Kaplan, a pro­fessor of physics at the University of California, Los Angeles. He had proven his organizational abilities by cofounding the university’s Institute of Geophysics in 1944 and by campaigning successfully for degree pro­grams in sports. As chair of the U. S. National Committee for the IGY, he concealed the habitual tension that turned him into a five-cigar man dur­ing sporting events.

The first meeting of the U. S. National Committee for the IGY took place at the National Academy of Sciences on the afternoon of March 26, 1953. For a day and a half, the group struggled to define a program and budget. Frustration mounted. One scientist suggested that they should for­get the whole thing. Berkner, a consummate committee man, smoothed over such moments, and by the end of the twenty-seventh they had out­lined their aims. In the meantime, they had solicited thoughts from their colleagues around the country.

Ideas poured into the academy during April. Many of them would sound familiar today. Issues were raised that today’s scientists continue to address. Paul Siple wrote, “There is evidence that the earth is undergoing a significant climate change, advancing from cooler to warmer conditions… our knowledge is still imperfect as to the exact cause of climate changes.”

By May the list of subjects to be studied included geomagnetism, solid-earth investigations, atmospheric electricity, climatic change, geodet- ics, cosmic rays, ionospheric physics, high-altitude physics, and auroral physics.

It is hard to imagine any study of these subjects today that would not rely partially or wholly on satellite observations. Then satellites did not exist, and the National Committee of the IGY did not initially consider that satellites should be developed. It was, however, clear to them that they needed a strong program of sounding rockets to probe the upper atmo­sphere. Some scientists were concerned that the rockets would cost too much, and perhaps price dissuaded them from adding satellites to their agenda.

The estimated price tag for this unprecedented research proposal, between 1954 and its closure at the end of 1958, was $13 million—$2.5 million more than the National Science Foundation (NSF) was requesting as its total budget for fiscal year 1954. The NSF and the academy under­took the task of requesting the money from Congress.

Behind the scenes, Kaplan lobbied hard. There were times when the project seemed doomed. A scientist in the administration told him, “Joe, go home. Such a beautiful program does not stand a snowball’s chance in hell of getting support.” Yet, it did.

Through the end of 1953 and 1954, planning for the IGY went ahead in the U. S. Concurrently, support for launching a satellite, despite deep skepticism from many, was gaining momentum among a vocal and persuasive minority in the military, industry, and academia. Some saw the IGY as the natural home for a satellite program, and they set in motion events that led to the General Assembly of the IGY endorsing the inclu­sion of satellites in its program.

This was in Rome in the fall of 1954. The night before, Berkner and others who were to become leading space scientists had debated until the early hours whether their enthusiasm was getting ahead of their ability, and whether they should seek the approval of their international colleagues. Slowly the enthusiasts converted the cautious. As they argued and won their case the next day, the Soviets simply observed—in silence. It was October 4, three years to the day before the launch of Sputnik I.

The International Geophysical Year was now close to its decisive encounter with the Eisenhower administration’s national security policy. While Berkner and his colleagues prepared for the Rome meeting, a top – level scientific panel, authorized by President Eisenhower in July 1954, was in the process of assessing the technological options available to prevent a surprise attack on the U. S., particularly by the Soviet Myacheslav-4 inter­continental BISON bombers. This panel, headed by James Killian, a confi­dant of President Eisenhower and the president of the Massachusetts Insti­tute of Technology, separated its task into three areas: continental defense, striking power, and intelligence capabilities.

Their final report contained a recommendation that the administra­tion fund development of a scientific Earth satellite to establish the free­dom of space in international law and the right of overflight.

The report was finally completed in February 1955, and later that month, Donald Quarles, the assistant secretary of defense for research and development, discreetly asked some members of the U. S. National Com­mittee for the IGY to formally request a scientific satellite.[2]

In response to the Rome resolution, the national committee had already asked a rocketry panel, which met for the first time on January 22, 1955, to “ … study and report on the technical feasibility of the construc­tion of an extended rocket, from here on called the long-playing rocket, to be launched in connection with scientific activities during the Interna­tional Geophysical Year.” “Long-playing rocket” was a euphemism for satellite launcher, and the new panel was told that its discussions should remain private.

By early March the panel had concluded that a satellite launch was feasible within the time frame of the IGY, but that guidance would be dif­ficult. On March 9, the executive committee of the U. S. national commit­tee debated whether to accept the panels findings. After some heated discussion, the report was accepted and the national committee requested that a satellite program be included in the IGY.

Quarles took the request to the National Security Council, which accepted the IGY’s satellite project on May 26, 1955. The next day, Eisenhower, too, endorsed the project. On July 29, President Eisenhower told the world that the U. S. would launch a satellite as part of the IGY. Thanks to Eisenhowers national security aims, the scientists who had campaigned for science satellites had what they wanted. Many of them, of course, did not know of the underlying agenda that had furthered their aims.

Within days of Eisenhowers announcement, the Soviet Union, by then a participant in the IGY, talked openly about its satellite plans at a meeting of the International Astronautical Federation in Copenhagen.

Leonid Sedov, an academician and chair of the impressively named Commission for the Coordination of Interplanetary Flight, made the announcement. He dropped into one of the sessions and, through his interpreter, called a press conference at the Soviet embassy During the conference, Sedov said that the Soviet Union planned to launch a satellite in about eighteen months, six months earlier than the earliest American estimate. The plans he outlined were for a much larger satellite than those the U. S. was planning. For those whose job it was to analyze Soviet inten­tions, here was public confirmation of the heavy-lift launch capabilities that the Soviets aspired to and a clear announcement that missiles capable of intercontinental distances were in the offing.

Sedov’s prediction of the Soviet launch date was wrong by ten months. But the Soviets did launch first, and their satellite was bigger than anything American scientists really believed in their hearts that the Soviets were capable of.

Chapter ten: The Realities of Space Exploration

The account of the launch of Transit 1A is pieced together from com­ments from different team members. Lee DuBois, for example, remem­bered the tears in his eyes when the satellite failed (page 105).

Details of the twenty-five-minute flight and the results gleaned come from the records that Henry Elliott had kept and from reports in APL’s archives (page 106).

Numerous memos and reports in the APL archives testify to Kershner’s industry in preparing for Transit IB and Transit 2A (pages 106 and 107).

John Hamblen’s undated, typed note requesting the team members to document component testing (page 107) is among the papers in the APL archives.

My description of how the launch might have been is pieced together from people’s memories and photographs of later launches found at APL.

A progress report details what happened scientifically following the suc­cessful launch of Transit IB (page 108- 109). Bill Guier explained what the report meant and supplemented it with his own memories.

A textbook consulted on the geoid (page 110) is Theory of the Earth, by Don L. Anderson (Blackwell Scientific Publications, 1989).

Information about how APL’s understanding of the geoid developed comes from papers and from interviews with Bill Guier and Harold Black (page 112).

Dave Smith, director of the division of terrestrial physics at the Goddard Space Flight Center gave me some very basic understanding of satellite geodesy.

Information about developing subsequent orbital determination programs and about the computers comes from reports in APL’s archives and from interviews with Harold Black and Lee Pryor (pages 113 and 114).

Information about the problems with the solar cells (page 114) came from Bob Danchik.