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.

New Moon

“In Leningrad, Korolev could not then by any means know that, after many very hard times, sometimes cruelly unjust to him, a beautiful spring would come when… would be reflected a world of black sky and blue Earth, a world never before seen by Man/*

—Yaroslav Golovanov, Sergei Korolev:The Apprenticeship of a Space Pioneer

“And all of a sudden you wake up one morning, and here’s this doggone Russian thing flying overhead… Oh, no, there was a great deal of disturbance..

—William Pickering, director of the Jet Propulsion Laboratory, in 1957.

From a transcript of an oral history in the archives of the California Institute of Technology.

I

f they had known then what they know now, would they have done the job? They would surely have been daunted, even given the imperatives of the Cold War. But they did not know how difficult space exploration would be. The “cold warriors” had their incentives: intercontinental ballis­tic missiles and reconnaissance satellites. And the enthusiasts, who some­times were also cold warriors, had their long-held aim: to go beyond Earth’s atmosphere. By Thursday evening, October 3, 1957, their destina­tion was less than a day away.

That same evening was one of the last on which delegates gathered at a conference organized by the National Academy of Sciences in Washington D. C. as part of the International Geophysical Year. The focus of the con­ference was on the rockets and satellites to be launched before the end of 1958. The IGY united sixty-seven nations in a seemingly impressive pax academica, yet, despite good intentions, the satellites of the IGY were about to push the world into a new phase of the Cold War.

Satellites were not part of the original plan. Indeed, when the IGY first endorsed satellites and chose as its logo a satellite orbiting the earth, many considered satellites to be little more than science fiction. In three
years that had all changed. By October 1957, the U. S. and USSR were fol­lowing one another’s progress keenly Each group of scientists wanted to be the first, and the meeting at the National Academy of Sciences gave them all an opportunity to probe each other’s intentions.

The Soviets were not very specific. Their delegation, led by Lieu­tenant General Anatoly Blagonravov, knew that a launch was imminent and had announced this on the first day of the conference. But they had not said exactly when. It is doubtful that they knew.

The Soviets would be gratified if the launch took place during the meeting, but they also knew, as the defector George Tokady was to say years later, that “ … the launch of Sputnik was too big a piece of cake to play games with.” Already the centennial celebrations on the birth of Kon­stantin Tsiolkovsky, Russia’s “father of spaceflight,” had passed on Septem­ber 17th without a satellite attaining orbit. The cognoscenti had speculated that the seventeenth might be the occasion for a Soviet launch.

But the satellite would be launched when all was ready, and one man would make that determination. Late Thursday afternoon, as the confer­ence workshop on rocketry struggled with the usual committee minutiae, that man, Sergei Pavlovich Korolev, lay—perhaps—restlessly in bed. More likely, being who and what he was, he paced the sitting room of his small cottage. It would be nearly a decade before the Soviet leadership publicly acknowledged the existence of this man, whom they called the Chief Designer of Cosmic-Rocket Systems.

Korolev’s cottage was on the grounds of Baikonur Cosmodrome, near the village of Tyuratam in Kazakhstan, one hundred miles east of the Aral Sea. It lay on a parallel with northern Wisconsin and as far east of the Greenwich meridian as Halifax, Nova Scotia, is west.

In Washington D. C., as October 3 drew to a close, workshop coordinators pulled ideas together for the next day.

For Korolev, it was the early hours of October 4. That day Korolev’s team would open the space age.

In future years, on the nights before he sent cosmonauts into space, Korolev would sleep very little. Instead, he worried about the well-being of the young men whom he drove as hard as he drove himself, buoyed by an energy that made his days so much longer than those of other people. But his wakefulness had a second purpose. It helped him avoid dreams in which guards beat him and screamed at him, dreams from which he would wake to a visceral fear of annihilation. These were Stalins legacy, as were his memories of colleagues who one moment were working alongside him and the next were gone.

On October 4, as the sun rose on a clear, cool dawn, Korolev was— comparatively—free. On his wall hung a portrait of Tsiolkovsky. Tsiol – kovsky had died twenty-two years before, but he had been the first to write scientifically about this day. His books, calculations, and ideas had long ago seduced Korolev.

Less than a kilometer away, beyond the small knot of trees that sur­rounded his house, the rocket glinted in sunlight. Beyond that, where only two years before no launch site had been, the semiarid steppes of Kazakh­stan stretched to the horizon and beyond. The American government had known of this launch site, built with forced labor, since the spring of 1957, when a U2 spy plane had returned with photographs.

For Korolev there must have been tension, anticipation, excitement, and fear as he faced the day that would test all that he had become and all that he had dreamed.

As Korolev prepared to leave for the launch site, perhaps he thought briefly of the people waiting for him to fail, the naysayers and his rivals. Earlier in the year, as rocket after rocket had exploded and Nikita Khrushchevs impatience had mounted, Korolevs detractors had pushed for his dismissal. Then, in early August, his team had successfully launched the R7, the worlds first intercontinental ballistic missile. Not for eighteen months would the U. S. launch a missile capable of similar range. A se­cond successful launch followed, and on August 27 the Soviet Union announced that it possessed intercontinental ballistic missiles. It would have been more accurate to say that they had the beginnings of the capa­bility, but the success had satisfied Khrushchev, who now had a rocket that would eventually be capable of carrying a two-ton thermonuclear warhead to the heart of America. As a result, Korolev had won the final go-ahead for his dream, the opportunity to send an artificial moon into space.

The following two months were full of frantic activity. Korolevs team had immediately begun intense preparations for a launch, and toward the end of August the satellite was ready to ship to the Cosmodrome. Korolev moved to his cottage to supervise launch operations. In Septem­ber, the pace picked up and tempers frayed. The satellite developed an electrical fault. Everyone panicked. Korolev, always relentlessly demanding, became merciless. Time and again, the launch team watched with appre­hension as his little finger rose to stroke his eyebrow; it was the signal to move smartly to the next job. They knew his capacity for compassion and for explosions of wrath; that he found people who would argue a case interesting, but would take personal offense at anyone who did not do his job. Korolev drove his engineers and his engineers drove themselves until, crisis by crisis, they coaxed the novel technology to readiness.

At last, toward the end of September, the crane in the assembly building hoisted the small, shiny sphere into the nose cone of the rocket. The launch team was now ready to put on a show for VIPs from Moscow. Members of the State Commission, the secret group that was to control the country’s space program, flew in. Technicians ran a final check on the satellite’s radio transmitter, switching the signal to a loudspeaker so that the commissioners could hear it echo around the building. Then the engineers silenced the transmitter. Some recalled later that their skin tin­gled at the thought that the satellite would not speak again until it was in space.

On the night of October 2, the launch team moved the rocket. It was four stories high and weighed nearly three hundred tons. Slowly, so very slowly, they wheeled the rocket out of the assembly building on a flatcar, and it began its painstakingly careful journey down the railroad track to the launch pad. It swayed with each uneasy movement. The next morning, Thursday October 3, they began the final preparations for launch—the countdown.

As the day progressed, the launch team worried that the satellite would overheat despite the gaseous nitrogen circulating inside the sphere. They threw a white blanket over the nose to give the satellite further pro­tection from the sunlight. Later, dissatisfied with that solution, they pumped compressed air around the nose cone.

On Korolev’s recommendation, the satellite ensconced in the nose cone was of a very simple design—a two-foot diameter sphere weighing 184 pounds. A sphere, Korolev said, was a fitting shape for what might be the world’s first satellite because it mimicked the shape of the natural bod­ies of the universe. Consummate engineer that he was, it seems that he could never quite suppress the poet in himself.

Soviet scientists planned both optical and radio tracking for their satellite, with the intention of learning what they could about the earth’s gravitational field and the density of the upper atmosphere. That week in

Washington, Soviet delegates were reemphasizing the frequencies that their first satellites would transmit. And the conference workshop on tracking resolved to establish stations capable of tracking the Soviet satel­lites, a resolve that, even as they made it, was too late.

The Soviets called their satellite Prosteyshiy Sputnik (meaning “sim­plest satellite”). Inevitably, the design and launch team had shortened that to “PS”; and Korolev’s staff, who referred to him informally as “SP” (for Sergei Pavlovich), used the initials interchangeably for the man and the satellite.

PS’s surface was buffed, as were those of the American satellites, so that it would shine in orbit as a sixth-magnitude star and could be tracked visually. It had four whiplike antennas that were pressed between the inside of the nose cone and the satellite’s surface. When, or if, Prosteyshiy Sputnik attained orbit, these antennas would relay radio signals at a frequency that every amateur radio operator in the world would be able to detect.

Prosteyshiy Sputnik’s destination was not far away, for the boundary of space is less distant from Earth than New York is from Washington D. C. Yet every inch of that journey would be fought against Earth’s overwhelm­ing gravity, which would yield only reluctantly to human ingenuity. If the launch vehicle did not reach a high enough velocity, PS’s trajectory would be a ballistic path high through the atmosphere and back to the Earth’s sur­face, like that of the ICBM that Korolev had launched in August. Alterna­tively, PS might be released into too low an orbit and burn up in the atmosphere.

Prosteyshiy Sputnik’s launcher was designed to put a far heavier cargo in space, but Korolev was moving conservatively. The heart of the launcher was the R7 ICBM; four cone-shaped, strap-on boosters surrounded its base, resembling a stiff pleated skirt. The boosters would peel away when their fuel was spent, leaving the R7, minus the burden of the boosters’ weight, to make the final push to orbital velocity. At liftoff, the R7 and the boosters, each containing a cluster of engines, would push from the Earth with more than half a million kilograms of thrust: power enough, if the launch was successful, to boost Nikita Khrushchev’s domestic reputation and help him to ward off those remaining critics who had participated in a failed attempt to oust him from power just a few months earlier. Khrushchev’s improved status should, in turn, help Korolev win backing for a continuing space program. A failure would set back Korolev’s aspira­tions, for space was not Khrushchev’s dream.

As it turned out, the wave of excitement that was to sweep the globe when Sputnik was launched seemed to take Khrushchev—and President Eisenhower—unawares. Prosteyshiy Sputnik, the simplest satellite, changed completely the world public’s perception of the Soviet Union’s technical capabilities, and Khrushchev learned quickly that space could yield politi­cal advantages. As a result, Korolev’s aspirations were to be harnessed tightly to Khrushchev’s international political goals, often in ways that cost lives and held back scientific advances.

On the morning of October 4, Korolev knew none of this. He needed success. But Korolev was also a dreamer, on a grand and generous scale, and he had dreamed this dream for thirty years. He had first worked with rockets after graduation from the Moscow Higher Technical School. He had stood on a sidewalk in Moscow as Friedreich Tsander, a fellow dreamer and space pioneer, raised his fist and said,“Forward to Mars!” That was in 1930, after the two young men had spent the evening with friends planning a research group to develop rockets and rocket-assisted aircraft. They called themselves the Group for the Study of Reaction Propulsion (GIRD).

Korolev regarded Tsander, who talked of rockets as though they already existed, as an older brother. At first, the two men and their friends had worked with no official backing or financial support in the cellar of an abandoned warehouse. They soon attracted the attention of the Soviet armaments minister, Mikhail Tukhachevskiy, a stroke of good fortune that won the group financial success but within a few years would lead to tragedy. In the meantime, GIRD expanded and joined the Gas Dynamics Laboratory in Leningrad to form the Rocket Research Institute. Korolev was appointed the deputy director, responsible to Tukhachevskiy

Before GIRD moved to Leningrad, they designed, built, and tested the Soviet Union’s first liquid-fuelled rocket. Korolev lit the fuse. The rocket, which was based on many of Tsander’s ideas, flew successfully on August 17, 1933, and landed 164 yards from the launch site.[3] Sadly Tsander, who had died in March at the age of 46, did not witness the short flight.

Despite what must have been grief at the death of his friend, petty restrictions limiting access to foreign journals, and the scarcity of food, even with ration cards, those were good years for Korolev. The govern­ment funded the institute because rocket research conformed to the national political goal of establishing Soviet technical supremacy In Leningrad, Korolev met Valentin Glushko, another Soviet space pioneer who designed rocket engines. In future years, Chief Designer Korolev was to collaborate often with Glushko, though the two men were to develop a stormy relationship. At the same time, Korolev’s personal life expanded. He married his school sweetheart, Xenia Vincentini, a surgeon, and they had a daughter, whom they named Natalia. But when Natalia was three, every­thing changed.

In the early hours of June 27, 1938, Stalin’s secret police, the NKVD, arrested Korolev. Though he did not know it at that moment, this was the end of his marriage and the beginning of torture, hunger, and years of imprisonment. He was charged with anti-Soviet activity, and his guilt was determined apparently by his association with Mikhail Tukhachevsky. Tukhachevsky was already dead, condemned and shot a year earlier on the strength of false documents that some suspect had been planted by the Nazis.

The NKVD packed Korolev into a boxcar of the Trans-Siberian Railway destined for Magadan. From there he was transported in the hold of a prison ship to the gold mines of Kolyma, concentration camps where thousands died each month. For nearly a year he was hungry, far hungrier than in the ration-book days of the early thirties. He lost his teeth, devel­oped scurvy, and in winter often woke to find his clothes frozen to the floor; but he survived.

He survived because the authorities transferred him to Moscow, to another kind of prison, one that held the cream of the Soviet Union’s aeronautical designers. Andrei Tupolev, the country’s most eminent aircraft designer at that time, headed the technical work of these scientists and engineers, though he was himself a prisoner.

When new prisoners arrived from the Gulag, Tupolev would ask them for a list of the engineers whom they had left behind in the camps. Some were reluctant to make such a list in case their colleagues had been freed and would be rearrested. That was probably not Tupolev’s intention, and he may have seen Korolev’s name on such a list and asked for him to be transferred to Moscow Tupolev would have recognized the name because he had supervised Korolev’s diploma project—the design of a two-seater glider—during Korolev’s final year at the Moscow Higher Technical School.

Whatever the reason for the transfer, Korolev found himself working long hours in a Moscow prison with sparse comforts. The authorities had a twofold work incentive scheme. They held out the hope of eventual freedom, and they threatened the prisoner’s families. Compared to Kolyma, it was paradise. But now Korolev knew that life could change any moment at the whim of a faceless bureaucrat. He tried repeatedly to impress this knowledge on fellow prisoners—that they might disappear without trace and no one would know of it. That knowledge was the darkness that would follow Korolev through his life, the darkness he would share with his friends in future years when late-night conversation lasted into the early morning hours.

Korolev spent his days like the other prisoners, designing aircraft for the war effort; at night, in the communal dormitory, he worked on rock­etry. With the end of World War II, his rocketry research once more emerged into daylight. The USSR and the U. S. were vying for men and equipment from Peenemiinde, the base where Germany had developed the rockets that bombarded London, Paris, and Antwerp in a new kind of warfare. As the Red Army swept into eastern Germany, Stalin remembered the rocketeers he had imprisoned eight years earlier. Back then a magis­trate had told Korolev, “We don’t need your fireworks and firecrackers. They are for destroying our leader, are they not?” By 1945, Stalin had changed his mind, and he turned to those rocketeers who had survived his purges.

Stalin was far more strongly committed to missile development than was the U. S. leadership at that time. In 1945, Stalin insisted on seeing pris­oner Korolev. Korolev always remembered how without taking his pipe from his mouth Stalin had demanded information about the potential speed of missiles, their range, payload, and accuracy. In his memoirs Khrushchev says that the rationale for official interest was that the U. S. could, if it wanted, station long-range bombers at air bases in Europe close to the Russian border, whereas the Soviet Air Force could not reach the continental U. S.

Korolev was sent under guard to Germany to glean what he could. He was under orders to track down those V2s that were not on their way to America and to select German engineers from among those who had not surrendered to American troops. Korolev sent both rockets and engi­neers to Russia. In the years immediately after World War II, both the Soviet Union and the U. S. learned as much as they could from German advances in rocketry while simultaneously developing their own missiles. In Russia, the Germans worked separately and did not know what Soviet engineers were doing, which disappointed western intelligence workers when Stalin sent the Germans back to Germany in 1951. Yet this should not have surprised them. Von Braun, who worked on intermediate range ballistic missiles in the U. S., could likewise not have known details of the Americans’ work on ICBMs.

By 1951 Korolev was no longer a political prisoner, at least not obvi­ously so. He and Xenia had divorced in 1946, and he had since remarried. He worked fanatically hard, prompting colleagues to wonder whether he had a home life. Of prison he rarely spoke. Occasionally, he would sip cognac late at night and reminisce with other former prisoners, telling them how those days still haunted him in his dreams. And a few nights before he died,[4] he told two close friends, the cosmonauts Yuri Gagarin and Alexei Leonov, of some of the pain. His death on an operating table in 1966 was, some speculate, a result of poor health stemming from his days in Kolyma.

In many ways Korolev’s early life must have prepared him for priva­tion.[5] He was born in Zhitomir, in Ukraine, in 1906. Three years later his parents separated, and Sergei went to live with his maternal grandparents while his mother went back to college. She left instructions that the child was not to leave the garden to play because she was afraid his father, who had already threatened her with a pistol, might kidnap him.

Korolev remembered those years as lonely ones. For a few years he wrote poetry. But he remembered vividly that when he was six, his grand­mother who was fascinated by gadgets and technology took him to see his first aircraft. This experience touched Korolev as deeply as seeing ballet or theater might define the destiny of another human being.

On weekends his mother would visit. Then they would play, and she would read to him, but she also made him go alone to distant dark rooms because, inspired by her reading of James Fenimore Cooper, she thought that in this way he would conquer fear.

As they did to millions of others, World War I and the Russian Revo­lution turned Korolev’s life upside down. In 1917, Sergei was eleven. By then his mother was married again, this time to an engineer called Grigory Balanin. The family lived in Odessa, on the Black Sea, and Grigory and Sergei settled to the prickly business of getting to know one another. The world war and foreign occupying forces depleted food stocks. Civil war followed world war. The family was often hungry, and Korolev would hike with his mother to the countryside to barter for potatoes.

When the civil war ended, with Lenin in charge, Sergei’s mother and Balanin sent Korolev to the First Construction School in Odessa. Here he learned physics and mathematics (one of the school’s mottos was “mathe­matics is the key to everything”) and how to tile roofs. He soon decided that he did not want to be a roofer, because by now he was fascinated by the idea of flying and designing aircraft and gliders.

After an obligatory summer as a mediocre roofer, Korolev applied to and was accepted by the Kiev Technical School. Two years later, he moved to the Moscow Higher Technical School where he earned his diploma in aeronautical engineering. A year later Korolev graduated from flying school. In the meantime he earned money by working in an aircraft design bureau.

He was now twenty-four, interested in rockets, but passionate about designing and flying sophisticated, record-breaking gliders. And in this role, anyone who cared to watch would have seen the character of the future chief designer emerge. Oleg Antonov, another of Russia’s great air­craft designers and also a gliding fanatic, met Korolev at about this time. Korolev was attempting a record flight in a glider of his own design. Inad­vertently, Antonov sent Korolev aloft with the anchor still attached to his glider. Korolev flew for four hours nineteen minutes, oblivious of the anchor. When he landed and saw the hole in the tail of his glider, he offered to tear Antonov’s eyes out with a pair of pliers. Yet Antonov remembered Korolev as a man of iron will and boundless humor.

By the morning of October 4, 1957, that combination of will and humor had carried Korolev to the position of chief designer of rocket – cosmic systems. Stalin had been dead for four years. After Stalin’s death, Korolev had been invited to join the Communist Party and had been elected a corresponding member of the USSR’s Academy of Sciences.

Ostensibly, he was a secure member of the establishment, yet the injustice of wrongful imprisonment ate at Korolev. He had asked repeatedly, and without success, to be rehabilitated—for his conviction at Stalin’s hands to be rescinded. Only after the successful launch of the satellite now sitting in the nose of the rocket on the launch pad would this happen.

So, on the morning of October 4, 1957, it was both the chief designer and the prisoner who awakened in terror, who turned a collar to the cold, climbed into his car, and drove to the concrete apron of the launch site. To the colleagues waiting for him, he was SP, a man rapidly becoming a legend among the few who knew of him. He was short and heavyset. He reminded them of a boxer or a wrestler, as much because of his personality as because of his physique. His brown eyes were bright with intelligence and passion.

Korolev saw himself as the designer whose role was to define the job, to listen to the team members, and then to make the decision. This he had done, paying attention to strategy and detail, balancing the consequence of one technical choice against another, seeking the right compromise. Now the product of hundreds of people’s work, of thousands of calculations, of designs and redesigns, was sitting on the launch pad. Would it open a new frontier?

Perhaps for propaganda reasons, perhaps because there would be an emotional symmetry in such an event, and perhaps because it is true, there is a persistent and disputed story that while at the Moscow Higher Techni­cal School, Korolev had visited Tsiolokovsky. The old man, so the story goes, told Korolev that rockets were a very difficult business, and Korolev had replied that he was not afraid of difficulties. It is hard to imagine that en route to the launch complex, Korlev did not remember Tsiolkovsky, who, when he died in 1935, was an old and nationally revered man. Per­haps, too, Korolev remembered Friedreich Tsander, his old friend from the free days in Moscow.

On October 4, 1957, Tsiokovsky’s, Tsanders, and Korolevs dream stood against the gantry, gleaming in the sunlight.

In Washington D. C., delegates awoke to the penultimate day of the rocket and satellite conference. A workshop was to debate what to include in the IGY’s manual on rockets and satellites.

New Moon

Korolev drove to the launch pad. The countdown was proceeding, and they were about to fuel the rocket. He mounted the platform to brief the engineers and to listen to accounts of the night’s doings. Then, while the launch team pumped liquid oxygen and kerosene into the tanks, he called Moscow with an updated report of the countdown.

Throughout the day, Korolev monitored everyone’s work, outwardly calm but fooling no one. By evening, technical difficulties had halted the countdown several times. Those who could stayed out of Korolev’s way The day moved inexorably forward. A day that for Korolev must have lain before him as a path to forever then passed in a second.

Thirty minutes before liftoff, everyone retired to their posts. Most went into a concrete bunker one kilometer from the launch pad. Those without a part to play in the final countdown climbed onto the bunker’s roof.

Korolev sat at a desk in the bunker, watching through a periscope. Floodlights bathed the rocket. Hope was palpable. As they waited and watched, someone walked beneath the floodlights. The unknown figure raised a bugle and blew clear notes into the midnight sky before hurrying back to safety.

Korolev listened as the loudspeakers relayed the deliberately spoken script of a space launch, a script that is still running, but which played that night to its first audience.[6]

“Duty crew, leave the pad.”

“Fire brigades, on alert.”

“Zero minus one minute.”

“Switch to start vents.”

Korolev knew that nitrogen was sweeping through the pipes, purging the giant rocket before oxidant and fuel met in a mighty chemical reaction. “Auxiliary engines pressurized.”

“Main engines pressurized.”

“Start.”

New Moon

Stillness enveloped the watchers. They dared not blink. And then it hap­pened. Incandescent vapors engulfed the rocket, throwing stark shadows on the surrounding concrete. The earth rumbled and a thunderous roar washed past their ears. They watched the huge rocket strain, and then—as if in slow motion—the engines lifted the rocket from the earth.

Did Korolev remember what Tsiolkovsky had written? “Mankind will not remain on the earth forever, but in the pursuit of light and space, we will, timidly at first, overcome the limits of the atmosphere and then conquer all the area around the sun.” Well it had begun. And what would those earlier versions of Korolev have thought of his fifty-one-year-old self, sitting, eyes glued to a periscope, watching his dream and an ancient dream of humanity’s ascend? Would the frozen wretch in Kolyma, with hunger griping in his belly, have believed this moment? What of the man who would start awake in terror or the child who saw with joy his first aircraft? What of the teenager in a civil war, the student, or the test pilot? Each gave a gift to the chief designer; surely they watched with wonder what together they had wrought?

Korolev slowly came to himself. Only minutes had passed, and already the rocket was a distant point of light. Around him people hugged and kissed, unshaven chins scraping cheeks a little damp. They danced and shouted, “Our baby’s off” Soon they fell silent and listened. The loud­speaker reported all systems nominal, later that the rocket had reached orbital velocity, and then that the satellite had separated from its rocket.

Now they faced another wait. Was the satellite in orbit? It should be overhead again in about ninety minutes. As the time approached that the satellite should be coming into range, they looked gravely at the radio operator. Then they heard the distinctive beep of their satellite in orbit, the signal they had last heard in the assembly building on that day a lifetime ago. The earth had a new moon. Sputnik I was in orbit.

Korolev notified Khrushchev and received the first secretary’s muted congratulations. The party machinery swung into operation. Soon the world’s teleprinters would carry news of the triumph and Soviet propa­ganda. Editors around the world would be galvanized as they read reports that included the words, “Artificial Earth satellites will prove the way for space travel, and it seems that the present generation will witness how the freed and conscious labor of the people of the new socialist society turns even the most daring of Man’s dreams into reality.”

At the cosmodrome, Korolev returned to his engineers. The chief designer, who had already experienced tragedy, now knew triumph. He was to know both again. That night, he mounted a platform and thanked his staff, those present and those at home. He was radiant. He continued, “Today we have witnessed the realization of a dream nurtured by some of the finest men who ever lived, including our own Konstantin Eduardovich Tsiolkovsky. Tsiolkovsky foretold that mankind would not forever remain on the earth. The sputnik is the first confirmation of his prophesy. The conquering of space has begun. We can be proud that it was begun by our country. A hearty Russian thanks to all.”

The Players

Direct evidence of field strength above the earth’s atmosphere could be obtained byV2 rocket technique, and it is to be hoped that someone will do something about this soon as there must be quite a surplus stock somewhere.

—Extra terrestrial relays: can rocket stations give world-wide radio cov­erage, by Arthur C. Clarke. Published in Wireless World

O

n October 4, 1957, only thirty-six people in the United States could call Europe simultaneously, via AT&T’s recently installed transat­lantic submarine cable—TAT-1. If the ionosphere was stable that day, about a further one hundred high-frequency radio circuits would have been available.

AT&T laid TAT-1 in 1956. It was a power-hungry coaxial cable, costing $2 million. To provide enough bandwidth (a wide enough range of frequencies) for live television would have needed twenty such cables. It was not until the mid 1970s that live cable TV was theoretically possible. By then satellites already spanned the oceans, though submarine cables made from optical fibers would mount a stiff challenge to satellites in the 1980s. But that, as they say, is another story.

So, in 1957, two or three people per state could have called Europe at the same time; even fewer could have called countries on the Pacific rim; there was no live transoceanic TV; and the information superhighway was an idea beyond even the most exotic pipe dream. Though computers, televi­sion, and telephone all existed, the oceans were truly barriers to communi­cation. And the world that contained these familiar-sounding technologies was very different from our own. After October 4, satellites, too, became a reality. Within decades, communications satellites had done much to change the world. Satellite communication is now a multibillion dollar business. Where did the story begin?

Before there were spacecraft, there were science fiction writers. Most imagined that ground controllers would communicate via radio with their spacecraft. Then in 1945 a junior officer in the Royal Air Force spotted the unique advantages for communication of putting a satellite into an orbit where it maintained the same position with respect to its subsatellite

point—geostationary orbit. The satellite would be like a huge microwave tower. Any antenna on Earth within sight of the satellite could beam a sig­nal to it, which the satellite would then amplify before beaming it back to another antenna at a different spot on Earth.

For a satellite to seem to remain stationary, it must meet two condi­tions: the orbit must take the same time to complete as the Earth takes to rotate once around its axis (be geosynchronous), and the plane of its orbit must coincide with the plane through the equator (zero inclination). If a geosynchronous satellite has an inclination of zero degrees, it is geostation­ary, and its place in orbit is designated by the longitude of its subsatellite point on the equator. A satellite is travelling at the velocity needed—more or less—to maintain a geosynchronous orbit when it is at an altitude of about 22,300 statute miles. At such an altitude, the satellite is within site of nearly one third of the Earth’s surface, excluding the poles. In 1959, satel­lites in this type of orbit were referred to as a 24-hour rather than geosta­tionary satellites.

The junior officer in the RAF imagined that this orbiting telecom­munications relay station would carry a crew, which, though wrong, was not a strange thought given the future science fiction career of the young man—Arthur C. Clarke. Clarke ran through some calculations, and in July 1945, he sent an article on the subject to the magazine Wireless World.

The editors were reluctant to publish something that seemed to them like science fiction, and they balked at acceptance. By October, they had relented, and the article appeared in print. It talked of field strengths and transmitter power, of solar power, and of how little time the satellite would spend in the shadow of the earth; and it suggested the best positions in orbit to provide a global system.[14] Clarke’s predictions turned out to be prescient.

At the other side of the Atlantic, an electrical engineer at the Bell Tele­phone Laboratory, John Robinson Pierce, who knew nothing of Clarke’s article, spent his leisure hours writing short science fiction stories and his working hours immersed in the complexities of microwave communication. Later, it was Pierce who was largely responsible for persuading NASA to carry out communication experiments with the passive Echo spacecraft in August 1960. Like the Moon, which reflected military communications between the East coast and Hawaii, Echo, which acted as a huge mirror in the sky, bounced a signal across the U. S. The two-way Moon relay was oper­ational between 1956 and 1962 and was manned when the Moon was in radio sight of both stations, usually for three to eight hours at a time. Often, when ionospheric storms shut down the usual radio channels, the Moon provided the only link to and from Hawaii for several hours at a time. In 1953, Pierce first suggested that if an artificial reflecting surface could be launched, it could bounce radio signals across oceans.[15]

Pierce joined Bell Laboratories after being awarded his doctorate by the California Institute of Technology in 1936. During World War II, he came across publications by an Austrian refugee, Rudi Kompfner, who was working for the British Admiralty. In 1943, Kompfner invented a class of vacuum tube, known as a traveling wave tube (TWT), that was to have an enormous impact on missile guidance and on communication through submarine cables and via satellite. In 1945, Pierce wrote his first paper about the new concept of traveling wave tubes and developed the first practical application of the technology. Kompfner would later say that he had invented the traveling wave tube but that Pierce had discovered it.

Pierce may as well be allowed to define a TWT in his own words, written in 1990 for a Scientific American publication, Signals, the Science of Telecommunication.

The traveling wave tube is a type of vacuum tube that gives high gain over a broad band of frequencies. An electromagnetic signal wave travels along a spring-shaped coil of wire, or helix, while electrons in the high voltage beam travel through the helix at close to the speed of the signal wave. The electrons transfer power to the wave, which grows rapidly in power as it travels down the helix.

Essentially, the vacuum tube allows electrons to flow from cathode to anode with few collisions and permits an energy exchange from the elec­tron beam to a radio wave constrained to travel the length of the tube. Thus the radio wave is amplified. There have been many versions of TWTs since 1945.

Pierce was impressed by what he knew of Kompfner’s work, and after the war he encouraged senior staff at Bell to recruit Kompfner. They were eventually successful, and Kompfner joined the lab in 1951. Pierce and Kompfner worked together cooperatively and productively for many years, and Kompfner was supportive of Pierces interest in satellites. Their work was the starting point for the Bell team that designed a second type of communications satellite, one with an active repeater that would, like the satellites envisaged by Clarke, amplify the signal before radiating it to Earth. This satellite became known as Telstar.

Telstar was not a twenty-four-hour satellite but rather was planned for a medium-altitude orbit, and so could only be seen by two ground sta­tions simultaneously for about twenty minutes. AT&T calculated that about forty 150-pound satellites in random medium-altitude orbits could provide a communications system with Europe. As soon as one satellite disappeared over the horizon, the transmitting and receiving antennas would lock on to the next mutually visible spacecraft. Such a system, said AT&T in the spring of 1961, could provide sixty channels by 1963 from North America to Europe and three thousand by 1980. The system would give ninety-nine percent probability of a satellite being simultaneously within sight of ground stations located in Maine and in Brittany.

In the summer of 1959, while the lab was still working on Echo, Pierce, Kompfner, and their colleagues at Bell were beginning to think that medium-altitude active repeaters rather than passive satellites were the most promising technology for transatlantic communication. For sound technical reasons, it seemed to them that geostationary satellites would not be feasible for many years.

On the West Coast, unbeknownst to Bell, a handful of engineers— Harold Rosen, Donald Williams, Tom Hudspeth and John Mendel— would soon solve, at least on paper, the problems then facing engineers considering a geostationary orbit. When Leroy Tillotson, at Bell, finished a technical paper on the specifications for a medium-altitude satellite in August 1959 and sent it to Pierce, Kompfner, and other senior members of the research department, Rosen and Williams were putting the finishing touches to their proposal for a twenty-four-hour satellite. The lightweight TWT designed by John Mendel, who had learned his trade in John Pierce’s lab, was critical to the proposal. The proposal, in Rosen’s and Williams’s names, was the beginning of a development that led to Syncom, the first geosynchronous satellite, and to Early Bird, the first commercial communications satellite.

In Arthur C. Clarke’s view, Pierce and Rosen are the fathers of com­munications satellites. During the early 1960s, however, there was little love lost between the two men. Rosen saw Pierce as obstructionist; Pierce thought that Rosen was making wild claims and would say anything to win support for his twenty-four-hour satellite. It is said they almost came to blows on stage during one conference. Yet they had far more in com­mon than either could have imagined. Both were told at different times by their superiors that they could not go ahead with their work. “Cease and desist,” is what Mervin Kelly, head of Bell Telephone Laboratories, told Pierce in 1958. Both had a fine disdain for the Department of Defense’s plans for a twenty-four-hour satellite and for NASA’s specifications for a medium-altitude active repeater called Relay. Both wanted to keep the government out of communications satellites.

Both, too, had been educated at Caltech, which in the 1930s was like an American Gottingen for the physicists and engineers who would become America’s leaders in aerospace. Pierce was a contemporary ofWilliam Pick­ering and sought his cooperation for the Echo experiments. Rosen was Pick­ering’s student and says that he was one of the teachers from whom he learned most. Each left Caltech in little doubt of his own intellectual ability.

The players, then, were John Pierce, Harold Rosen, and Donald Williams, with Tom Hudspeth, Rudi Kompfner, and John Mendel in strong supporting roles. Passive, medium-altitude active, and twenty-four – hour active satellites were the engineering concepts they contemplated.

Men and ideas fitted into a larger, more complex tapestry. It was not just that communications satellites were now within the state of the art, but there was also an increasing commercial and military demand for bet­ter communications.

TAT-1 remained the only transatlantic cable for telephony until AT&T laid a second link in 1959, bringing the number of simultaneous calls that cable could carry across the Atlantic to seventy-two. Adding these to the number of high-frequency radio circuits available on a good day, as many as four people per state could simultaneously have called Europe on the day of John Kennedy’s inauguration as president in January 1961. It was still not possible to make live transoceanic TV broadcasts. Instead, recordings were flown by jet or fed slowly down cables.

In response to growth in demand, particularly for calls to and from Europe, AT&T planned to lay a third transatlantic cable in 1963, adding a further two hundred telephone circuits. Even this would not be enough to meet predicted growth in demand. But there seemed to be insurmount­able engineering obstacles to developing higher-capacity cables, and the radio spectrum was already overcrowded. Worse still, solar minimum would occur between roughly 1962 and 1966. With less solar energy enveloping Earth, the ionosphere could be less active and thus would not reflect certain frequencies. Experts calculated that this would cut by two – thirds the available high-frequency radio channels worldwide.

The Department of Defense, with troops stationed around the world, often in places with which it was difficult to communicate, was deeply concerned.

With such a paucity of communications infrastructure coupled with the commercial and defense advantages of enhancing communications, it is not surprising that the Kennedy administration placed a high priority on the development of communications satellites. Communications satellites (and meteorology satellites) figured in Kennedy’s famous moon speech of May 25, 1961.

And there were strategic advantages for the United States in devel­oping communications satellites. Communications technology looked as though it could serve as a versatile foreign policy tool that could extend American influence throughout the world. John Rubel, deputy director of defense research and engineering (DDR&E) and for a while the acting director, pointed out in a white paper written in April 1961 that countries newly emerged from colonialism were often reliant for communication on their former colonial powers. He cited the cases of Guinea and Nigeria, which had to go through France and England to communicate with one another. It would be of “incalculable value” in the battle for men’s minds, wrote Rubel, for the United States to maintain a lead in communications technology. “Many feel that the United States should support satellite – based telecommunications systems to achieve these aims, even though there were no immediate commercial advantages resulting therefrom.”

The DDR&E held the third highest civilian position at the Penta­gon, roughly on a par in some circumstances with the chairman of the Joint Chiefs of Staff. Thus Rubel was in a position of some influence. He had exerted that influence once, at the prompting of NASA administrator T. Keith Glennan, to change an agreement that NASA and the Defense

Department had made in November 1958, confining NASA to work on passive communications. It was an agreement that deeply frustrated NASA’s engineers at the Goddard Space Flight Center. A new agreement, formalized in August 1960, permitted NASA to work on active commu­nications satellites. Both parties observed a tacit understanding that NASA would work only on medium-altitude satellites, while Department of Defense developed twenty-four-hour satellites.

By April 1961, Rubel seems to have been feeling his way through a complex strategy that would also set aside this second agreement and per­mit NASA to develop Rosen’s twenty-four-hour satellite. This was neces­sary because the Defense Department’s own plans for a twenty-four-hour satellite, called Advent, were going disastrously wrong, but there would have been too much opposition to simply canceling the satellite and replacing it with Rosen’s. If, however, the agreement between NASA and Defense could be set aside, then NASA could place a contract for the Rosen proposal. The agreement was dropped.

By August 1961, NASA had placed a sole-source contract with the Hughes Aircraft Company for a twenty-four-hour satellite, and the Department of Defense was to make the Advent ground stations available. The idea was that the twenty-four-hour proposal, now called Syncom, would be a cheap interim satellite to meet military needs until Advent could be developed. A year later, Advent was canceled.

The Syncom decision was a sweet triumph for Harold Rosen and Donald Williams. Before placing a contract for Syncom, both NASA and the Defense Department had been dismissive of Rosen’s and Williams’ engineering concepts. The first Syncom satellites were transferred to the military, and in the mid 1960s these provided links to Southeast Asia in support of America’s growing presence in the region. Thirty years have passed, and Harold Rosen is not yet tired of telling people how the Army and Air Force rejected [his ideas] but within a few years had to rely on Syncom.

By funding the development of Syncom, launching Telstar (at AT&T’s expense) and developing Relay, NASA enabled two approaches to a global satellite communications system to be demonstrated. When the International Telecommunication Satellite Organization (Intelsat) was formed in 1964, it was not yet clear whether international communica­tions would be based on twenty-four-hour satellites or constellations of medium-altitude satellites. If one twenty-four-hour satellite operated suc­cessfully, however, Intelsat would know that it was well on the way to pro­viding a global system, whereas tens of Telstar (or Relay-like) satellites would have had to be launched to prove that a global communications sys­tem of medium-altitude satellites would work. Thus it was sensible to first launch one twenty-four-hour satellite, and the success of three Syncom satellites was encouraging.[16] The successful launch of Early Bird settled the question, and most commercial communications satellites today occupy geostationary orbits (the countries of the former Soviet Union use another orbit, one better suited to communications at high latitudes).[17]

Though the decision to “go geostationary” has been validated since 1965, the merits of the alternative technological approaches were still being debated in the early 1960s. The technical arguments were enmeshed in and obfuscated by a highly charged policy debate about the role of government versus private industry in the development of communications satellites.

The debate began during the closing months of the Eisenhower admin­istration, when T. Keith Glennan announced that NASA would facilitate the development of communications satellites by providing launch oppor­tunities for industry on a “cost reimbursable basis,” which meant that industry would pay for the launch, but not at a true commercial rate. Glennan, like President Eisenhower, believed that private industry should be involved in the development of communications satellites. At the time, the most aggressive private industry in this field was AT&T. By October 1960 it had started Project TSX, which became Project Telstar, and had begun spending millions. Senior NASA staff and the attorney general were leery of AT&T. The company already had a virtual monopoly on voice transmissions. Neither NASA nor the Justice Department wanted to make decisions that would exclude from the new field companies that were not starting from the strong position of an existing monopoly.

When President Kennedy took office and James Webb replaced T. Keith Glennan, the emphasis shifted somewhat to a concern about how much control the government should retain over the development ol communications satellites given their strategic importance. Webb, whom

AT&T viewed as anti-industry, said that he did not want to put AT&T up against the whole Soviet Union.

Engineers at Bell perceived that the debate had become truly heated in February 1961 when Lloyd Berkner, whose proposal of an International Geophysical Year ten years earlier had set so much in motion, said in a speech that communications satellites would be a billion-dollar business in ten to fifteen years. The newspapers picked up the comment. Congress took note, and the Justice Department quoted Berkner in submissions to the Federal Communications Commission and Congress. Berkner’s com­ments were used to bolster the argument that space communication was too big for one company. Though his prediction was to prove correct, Berkner modified his views shortly after making them known, making the not unfamiliar claim that the media had exaggerated them.

But verbal arabesques could not change the course of the debate. Berkner had tapped into some widespread and deeply felt issues: the previ­ous administration’s concern about extending AT&T’s monopoly; the cur­rent administration’s desire to have some control over the development of a technology with strategic implications for the military, for commerce, and as a foreign policy tool; the current administration’s concern that a private company should not represent the United States in negotiations for a global system; and industry’s objection to being excluded by monopoly power from a potentially lucrative new market.

By February 1962, the Kennedy administration had sent a communi­cations satellite bill to Congress. The bill set up a private company called Comsat under strict governmental control. Half the stock was offered to the general public and half to the common carriers. The Federal Commu­nications Commission was responsible for distributing stock fairly between the common carriers, including AT&T. Key members of Congress had their own ideas about the bill, but by the end of the summer, the Senate had passed it, 66 to 11, and the House by 354 to 9.

President Kennedy signed the Comsat Act on August 31, 1962. It was the death knell for Telstar, though the concept of medium-altitude satellites had not yet been abandoned. Comsat would be the driving force behind the formation of Intelsat and thus behind the “go geostationary” decision. In the larger world of national and international policy, it was surely the right decision. To the engineers at Bell and to John Pierce, the man who pioneered the idea of commercial communications satellites and developed some of the critical technologies, the Comsat decision was a bitter disappointment.

Подпись:
Подпись: Of Moons and Balloons

More than thirty years later, in his home in Palo Alto, John Pierce dis­poses concisely and precisely of questions about his pioneering days, tugging all the while at a bushy eyebrow. With his sloppy yellow Labrador retriever in attendance, Pierce reminisces politely about Echo and Telstar. Clearly, he has told the story many times, and he says,“I prefer to look for­wards rather than back.”

Asked to explain how a klystron works, he grows more animated. He becomes even more interested when writing down the names of mystery and science fiction writers he has not previously come across or talking of the Chinese poetry he translates, the haiku he writes, his admiration for Milton and Blake. Only when I asked him about the Comsat decision did passion flash with the sharpness of a disappointment almost, but not quite, forgotten. Pierce, a loyal son of “Ma Bell,” would not leave Bell Labs, so the Comsat decision that excluded AT&T from international communication via satellites excluded him personally from a field he had pioneered.

As Pierce talks, his movements and speech are like those captured on video in the early 1960s, when he was the executive director of research at the Bell Telephone Laboratories. They are characteristically incisive move­ments, suggestive of someone who does not suffer fools or pretenders gladly. For a while, he thought that Harold Rosen was a pretender. “I was wrong,” he says. He is less charitable about some of those he encountered at NASA headquarters during the Echo and Telstar days.

Pierce retired from Bell Labs in 1971 at sixty-one, an age when, as he says, he was still young enough to do something else. He joined the faculty of the California Institute of Technology for nine years, then moved briefly to the Jet Propulsion Laboratory as chief technologist. He is now visiting professor of music emeritus at Stanford, where he pursues an inter­est in the psychophysics of music—the relationship between acoustic stim­ulus and what we perceive internally. It is an interest he developed at Bell in his postsatellite days.

John Pierce’s interest in science started when he was very young and his mother read to him from “very unsuitable books.” Long before he could read, John could say words like electromotive force, even if he didn’t

quite know what they meant. “She was the mechanical member of our family,” recalled Pierce. She also seems to have had faith in Pierces mechanical ability, because when he and his friend Apollo built a glider, she went up with him, apparently unfazed by the earth flashing by beneath her feet. This, despite knowing that the first glider they had built had fallen apart as it taxied for takeoff. “I was crazy in those days,” says Pierce, “doing things with very little information. I call it gadgeteering.”

Pierce studied at Caltech, and after changing his major a few times, he settled for electrical engineering. He graduated in 1933, looked around at his Depression-era employment prospects, and decided he would be better off staying at Caltech. He gained his master’s in 1934 and his doc­torate in 1936. This time the world outside the ivory tower was less hostile to him, and Pierce got a job at Bell Laboratories.

He was told to work on vacuum tubes and left to get on with it, despite knowing next to nothing about the topic. This was typical of Bell Laboratories, where there was a lot of intellectual freedom to pursue research as well as the money to pay for it. Perhaps that accounts for the nobel prizes awarded to physicists at the lab.[18]

By the time World War II broke out, Pierce was expert in the basic theory and design of various classes of vacuum tube. He applied that knowledge during the war and learned a lot about electron optics and broadband amplification. Pierce contributed to the body of work that opened the spectrum above thirty megacycles, which before World War II was almost empty of artificial signals. Developing that technology was essential to the feasibility of communications satellites.

It was while undertaking a mathematical analysis of broadband amplification that Pierce came across Rudi Kompfner’s work on traveling wave tubes. He was impressed. He wrote to Kompfner in 1946, adding his voice personally to that of the management whom he had persuaded to recruit Kompfner.

After working on traveling wave tubes, Pierce and others at Bell Labs turned their attention to MASERs (Microwave Amplification by Stimulated Emission of Radiation). These devices generate or amplify microwaves. When they amplify a weak signal, they add little noise. It was the MASER at the heart of the ground antenna that made it possible to pick up the reflected signal from Echo, which was only a million-million-millionth of the ten kilowatt signal beamed to the satellite for reflection across the coun­try. The MASER improved the antenna’s sensitivity by a factor of one hun­dred compared with what Pierce had envisaged when he first speculated on the use of an Echo-like satellite for communication. And it was this MASER, protected from extraneous ground noise by a horn-shaped dish, with which Penzias and Wilson detected the cosmic background radiation.

So by 1954, many of the ideas and devices that were crucial for Bell’s satellite communication work existed. And it was about now that Pierce became the first of the pioneers of communications satellites, which came about because he wrote science fiction stories (under the pseudonym J. J. Coupling, a concept familiar to electrical engineers). As a science fiction fan and author, Pierce was asked to give a talk on the subject of his choice to the Princeton, New Jersey, branch of the Institute of Radio Engineers. He must have had an erudite audience, given that RCA and Princeton University were nearby

Over the years, Pierce had given talks about man in space, but he decided that for this audience he wanted a less fanciful subject. He began to wonder what role satellites could play in his own field of communica­tion. At the time, says Pierce, communications satellites were “in the air,” though it was a rarefied air. In 1952 he had written an article about inter­planetary communication and had concluded that it was easier to commu­nicate between the moon and Earth than across the United States. Now he did some quick calculations of the power needed for transmission to and from orbiting spacecraft and was surprised to discover that communica­tions satellites were feasible.

Pierce gave the talk, which was to form the basis of his pioneering ideas for communication satellites.

Professor Martin Summerfield told Pierce that he should publish his talk. So, in November 1953, Pierce sent an article to Jet Propulsion, the journal of the American Rocket Society, which published it in April 1954.

The paper proposed three types of communications satellite: a one- hundred-foot sphere that could reflect a signal; a hundred-foot mirror in a twenty-four-hour orbit; and an active repeater in a twenty-four-hour orbit. The latter two, while theoretically stationary with respect to the ground, would actually be affected by solar and lunar gravity and so would need steerable ground antennas and stabilization by remote control.

The first of the three options—a hundred-foot sphere—was to be Pierce’s inspiration for the Echo communication experiments in 1958. In 1954, shortly after Pierce’s article was published, the U. S. Navy began experimenting with the voice transmissions to and from the moon that became the moon relay. But the moon is not an ideal reflective surface; its roughness gives multiple echoes at different wavelengths. A smoothly reflecting artificial satellite would, Pierce knew, provide a much higher – quality passive relay.

In 1954, few believed that satellites would be launched. Undeterred by the common view, Pierce told his audience in Princeton and wrote in his paper that if one found a way to build and launch a satellite, two classes of problem would remain, celestial mechanics and microwave communica­tion. First, they would need to know where their satellite was and would be; then they would need to send and receive radio signals. All the satellite operators had to come to grips with celestial mechanics; some, like Transit, needed a very detailed understanding of the earth’s gravitational field and its impact on an orbit. Pierce’s paper acknowledged the problem but devoted more time to the issues of microwave communication: signal losses on passage from the satellite through the ionosphere and atmosphere to Earth (path losses); the diameter of transmitting and receiving antennas; signal frequency and strength; radio beam width; the method selected for superimposing the signal, such as voice or music, onto the radio carrier wave (modulation); the nature of the polarization of the radio beam; the frequency of the carrier radio wave; sources of noise (that is, other fre­quency sources that would make the signal difficult to hear); the power of the signal; the signal-to-noise ratio and the sensitivity of the receiver. These were among the topics that five hundred scientists and engineers would later address during the Telstar project.

The science fiction books that Pierce had begun reading as a teenager made spacecraft and radio communication commonplace ideas to him. So the topic of his talk to the Princeton radio engineers is not sur­prising. But his early love of science fiction also held him back. He had been so used to thinking of spacecraft as romantic fantasies that he did not at the time realize how close they were to realization in his own field of communication. Pierce discussed the idea of communications satellites with people around the lab, but he was concerned about the reliability of vacuum tubes (and who better to know their limitations) in space and the limited abilities of the primitive transistors that then existed. “I was conser­vative about satellites,” he says.

Nonetheless, Pierce was responsible for persuading NASA to con­duct communications experiments with Echo.

The satellite that became Echo was not initially intended to be a communications satellite. It was suggested by William O’Sullivan from the Langley Research Center when James Van Allen’s satellite panel was selecting experiments for the International Geophysical Year. O’Sullivan wanted to launch a giant, aluminized Mylar balloon that could be inflated to a diameter of one hundred feet. With its small mass and large surface area, the balloons would be sensitive to comparatively small changes in force and thus would allow scientists to record how atmospheric density varied with position and solar activity and affected Echo’s orbit. Van Allen’s panel thought the balloon would be a good idea if they had sufficient resources for more than four launches.

In spring of 1958, Pierce and Kompfner read about the balloon and realized that it was exactly what Pierce had imagined would make an ideal passive communications satellite. They packed an ohmmeter and went to Langley to measure the conductivity of the plastic balloon. They decided that it would have a high reflectivity for microwaves. They took some samples of the aluminum-coated mylar back to the lab with them and confirmed its reflectivity. All they needed now was someone to launch the balloon. Unfortunately, the balloon was not one of the high – priority experiments for the IGY. NASA had not yet been formed, and the Department of Defense was already thinking in terms of the elaborate satellite that eventually became Advent and was to go so drastically wrong.

That summer, Pierce and Kompfner went to a meeting on commu­nications satellites at Woods Hole. William Pickering (director of the Jet Propulsion Laboratory) was there and showed himself sympathetic to Pierce’s ideas. Pickering suggested to the meeting that O’Sullivan’s balloon would be ideal for a passive communication experiment. If Bell could find someone to launch the satellite, said Pickering, JPL’s Goldstone ground sta­tion would participate in coast-to-coast communication experiments. To

Pierce, it seemed that Pickering’s support was vital to the success of the lab’s subsequent discussions with NASA.

Pierce returned from the meeting to a mixed reception. Mervin Kelly, the president of Bell Telephone Laboratories, asked a mathematician to study Pierce’s proposal. The mathematician’s report was negative. Kelly told Pierce to “cease and desist.” Kompfner thought that their plans could go no further, but Pierce developed a severe case of deafness. He continued to think of Kelly as one of his heroes but concluded that “even great men” can be wrong. In October 1958, he delivered a paper on transoceanic commu­nication via passive satellites to a national symposium on extended-range and space communications.

Later that same month, Pierce served as a consultant to the Advanced Research Projects Agency’s ad hoc twenty-four-hour satellite committee. He listened to what he thought were impractical and inefficient proposals from “these completely uninformed men.” It was clear after this meeting, in which elaborate satellites were discussed, that the Department of Defense was not going to launch the hundred-foot balloon. And shortly afterwards the Defense Department and the newly formed NASA agreed that the Department of Defense would develop active satellites and NASA would develop passive ones.

In the meantime, William Pickering had remained interested in a communication experiment, and NASA had been born. The new agency immediately inherited the Langley Research Center and the Jet Propul­sion Laboratory.

In November 1958, T. Keith Glennan, NASA’s administrator; Hugh Dryden, the deputy administrator; and Abe Silverstein, the director of space flight development visited Bell. The purpose of the meeting was to discuss global communications problems. Pierce made a general presenta­tion about communications and satellites, and Kompfner talked about components, data processing, and tracking and guidance philosophy. The NASA contingent was interested, but nothing seems to have come of the discussions.

At the end of December, NASA took over Project Vanguard, thus gaining control of a launch vehicle. By January, the agency was showing an interest in BTL’s ideas. On January 22, NASAJPL, and Bell discussed what they hoped to learn from transmitting a signal between the East and West Coasts. Kompfner wrote to Leonard Jaffe, who headed NASA’s communi­cation satellite work, on February 10, urging him to let Bell know soon whether the project with JPL was to go ahead because of the large amount of work that had to be done. Less than a week later Kompfner warned all technical staff that a considerable amount of work of an unusual nature was coming up. Until now, Pierce had been deeply involved in selling the project to NASA. Now he took a back seat. The day-to-day running of the satellite work was handed over to Bill Jakes, who was responsible to Kompfner. At the end of February, Jakes was immersed in technical discus­sions with NASA about the MASER and how much bandwidth was needed given the signal and its Doppler shift. Pierce was already toying with the idea of active broadband satellite communication. NASA was by now enthusiastic and was contributing more money to Echos than was Bell. The lab was building the horn antenna for experiments.

On June 10, 1959, there was a large meeting of all those involved with the project. O’Sullivan reported that one full-size balloon, including its inflating mechanism, was already being tested.

The balloon was made of 0.0005-inch-thick Mylar, coated by a 2,000 angstrom layer of aluminum. It weighed 136 pounds and had an optical reflectivity of seventy-five percent for tracking and a radio reflec­tivity of ninety-eight percent. Being passive, it did not have the compli­cated electronics needed for active repeaters, but it did carry a radio bea­con so that it could be tracked by Minitrack. The sphere was to be inflated in one second by the release of four pounds of water through a plastic nozzle in the sphere. Langley calculated that the vapor pressure would last for seven days, and they were testing six subliming solids in an effort to extend the lifetime. After seven days, they expected a gradual loss of pres­sure because of micrometeorite impacts. The sphere would get wrinkled, decreasing its usefulness as a reflector for communications. Between November 1959 and July 1960, BTL and JPL practiced bouncing signals to one another first off the moon and then, three times, off TIROS. Their pointing accuracy needed to be good because Echo’s dimensions at an alti­tude of 1,000 nautical miles would be the equivalent of an object a little over an inch long a mile away, and it would be moving at four and a half miles per second.

The first launch attempt failed because the balloon did not inflate. But the second attempt, on August 12, 1960, was a success. Tracking Echo turned out to be tricky. The original plan was that NASA, at Goddard, would compute the orbital parameters and turn them into tracking in­structions for Goldstone and the Bell antenna. A. C. Dickieson, Transit’s project manager, writes in an unpublished manuscript that the tapes as received were late and full of errors. More success, he says, was achieved by taking orbital parameters generated by the Smithsonian Astrophysical Observatory and calculating tracking errors locally Errors in orbital pre­diction were, however, inevitable in the fall of 1960. Only a few months earlier Bill Guier had predicted Transits position in orbit and realized how much more complicated the earth’s gravitational field was going to be than anyone had thought.

From Bell’s perspective, Echo provided background information for system planning and the design of Earth stations—information that fed into the Telstar project. Echo also demonstrated the effectiveness of the lab’s low-noise receiving equipment as well as the predictability and stabil­ity of the transmission path.

Echo was the first satellite that was visible to the naked eye, and T. Keith Glennan had anticipated that it would cause a sensation. It did, bolstered by AT&T’s brilliantly executed publicity campaign. On the night of the launch, the company sponsored a news special on NBC. It was replete with portentous music and massive radio telescopes. Another AT&T-sponsored video opened with ‘America the Beautiful’ and tugged at the patriotic heartstrings. AT&T won the publicity stakes hands down, but alienated NASA. The company’s expropriation of Echo did not win it any friends at the agency. Pierce wonders whether AT&T’s publicity suc­cess with Echo influenced NASA’s selection of RCA for the Relay satellite. If it did, it was a minor influence compared with the much larger policy issues that were at stake.

Echo’s success, technically and with the public, encouraged AT&T to go ahead with the development of a medium-altitude satellite as a proto­type of a global system of communications satellites. That satellite— Telstar—became every bit as famous as its predecessor.

General comments on satellite navigation

1. At different places in this section, I have alluded to alternative ideas for navigation satellites. One, explained in “Navigation by Satellite” in Missiles and Rockets in October 1956, even talks of utilizing the Doppler shift for a navigation satellite. But this paper envisages almanacs and tables of posi­tion and calculations of the distance at closest approach. It implicitly assumes that the orbit would be known and, inevitably because it was written in 1956, does not account for the impact on orbits of Earth’s complex gravitational field, nor for the impact of the ionosphere on the received signal. The paper does envisage the use of computers, but not the sophisticated curve-fitting techniques of Guier and Weiffenbach.

2. “Possible Use of Syncom as a Navigation System—Microwave Loran.” Memo for files, From L. M. Field cc L. A. Hyland (НАС archives 1990-09 box 6 folder 22).

This memo argues that Syncom would make a better navigation satellite than Transit if the station keeping were adequate. It expands on a memo by Donald Williams (see communications section) written on September 1, 1959.

Cocktails and the Blues

“I think there is very little doubt that the Russians intend to start firing [satellites] sometime on or before the first of the year [1958].”

—Richard Porter, chairman of the U. S.Technical Panel on Earth Satellites,

October 3, 1957

“As it happened, the public outcry after Sputnik was earsplit­ting. No event since Pearl Harbor set off such repercussions in public life.”

—William McDougall,.. .The Heavens and the Earth, a Political History

of the Space Age

L

ieutenant General Anatoly Blagonravov sipped vodka. In Tyuratam it was the early hours of Saturday, October 5, 1957. But in Washington, D. C., it was still the evening of Friday, October 4, and Blagonravov was hosting a reception at the Soviet embassy for delegates to the IGY’s con­ference on rockets and satellites. Sputnik I was in orbit, had, in fact, already passed undetected over America. It would not again go unnoticed.

The people at the Soviet embassy did not yet know that the space age had begun. As the guests circled the buffet of Russian delicacies spread beneath sparkling chandeliers, the Americans probed Blagonravov for hints of when his country planned to launch a satellite.

Although Blagonravov headed the Soviet delegation, the Americans could not have known quite how ideally placed he was to answer them. He was an academician and chair of the Soviet Academy’s Interdepart­mental Commission of Interplanetary Communication; as such he knew what the Soviet scientists wanted to do with satellites. He was a former head of the rocketry and radar division of the Soviet Academy of Artillery Sciences, to which Sergei Korolev had been elected in 1954. Finally, Blagonravov was a member of the State Commission and one of the VIPs who had listened to Prosteyshiy Satellite’s test signal in the assembly build­ing at Tyuratam little more than a week earlier.

What no one who saw Blagonravov that night can say is whether he knew at the beginning of the reception that Sputnik was already aloft. The white-haired, scholarly figure mingled genially with his guests, his demeanor much the same as it had been all week. His blue eyes masked an inner intensity. His Russian cigarette was tilted between his lips.

What is indisputable is that Blagonravov knew that a launch was imminent. So did the American scientists—at least intellectually—includ­ing William Pickering, a native New Zealander of great charm and under­lying sharpness. In 1957, Pickering was a veteran rocket scientist and had been the director of the Jet Propulsion Laboratory in Pasadena, California, for three years. He worked closely with the German rocket pioneer Wern – her von Braun. These two knew that with comparative ease they could convert one of the intermediate range ballistic missiles (range 2,000 miles) they were developing into a satellite launcher.

Pickering had first become involved with rocketry in 1942, at the Jet Propulsion Laboratory (JPL). The lab had analyzed intelligence reports about a V2 rocket that had landed in Sweden. In 1943, Army Ordnance asked JPL to investigate the possibility of developing an American long – range rocket, which in those days meant rockets that could fly at least one hundred miles.

When the war ended, von Braun surrendered to American troops and was moved eventually to the Army Ballistic Missile Agency in Huntsville, Alabama. The U. S. Army also shipped home about one hun­dred V2s, which von Braun had designed. These rockets were distributed to groups around the U. S. that were interested in upper-atmosphere research and rocketry. A group of scientists known as the Upper Atmo­sphere Research Panel selected and analyzed the experiments that the rockets would carry Members of this group, including James Van Allen, went on to head the IGY s satellite efforts.

Pickering, in cooperation with the Army Ballistic Missile Agency and von Braun, was soon testing the V2s while JPL—known as the “army smarts”—simultaneously continued its own rocket development. At times, as rocket after rocket exploded, it seemed to Pickering that rocketry was an unpredictable art, not a science. But JPL and the Army persevered.

By the early 1950s, the V2s were proving the truth of what Sergei Korolev had said in 1934 in his book Rocket Flight in the Stratosphere: “ … a rocket is defense and science.”

Von Braun, too, had always known this. Like Korolev, he dreamed of space and worked on missiles. In his new country, von Braun was a highly visible advocate of space exploration. In the early 1950s, he drew headlines by promoting grandiose ideas of space stations and Martian colonies at a time when the general public and most scientists thought that even a sim­ple satellite was science fiction.

While the idea of a space station was indeed science fiction at that time, the technical pieces that would make the launch of a satellite a prac­tical project were in place in 1954. That September, von Braun and a small group of scientists drew up a plan for a satellite that would become famous in the annals of space history. His idea was that a Redstone rocket would lift a five-pound satellite from Earth and that an upper stage, comprising a cluster of Loki rockets atop the Redstone, would give the satellite its final kick into orbit. If the small satellite was successful, a second, larger, “bird” carrying scientific instruments would follow. This proposal was called Project Orbiter.

The result of the first shot would be five pounds of metal circling hundreds of miles above the earth at a speed of roughly 17,000 miles per hour, more than four miles per second. Impressive, but how could they prove that such a small, swift, and distant object was really in orbit?

Nowadays, of course, radio links satellites to observers on the ground. Then radio was not an obvious choice. Von Braun turned to astronomy. The satellite would, after all, be a new heavenly body.

No one was better qualified for the job of finding a body speeding through space than the man he turned to, Fred Whipple. Whipple was the director of the Smithsonian Astrophysical Observatory in Cambridge, Massachusetts. A tall, slender man with a widow’s peak and steel-rimmed glasses that gave him an air of erudition, he was expert at tracking meteors and comets. Whipple concluded that optical tracking could follow a bur­nished satellite at an altitude of two hundred miles, even though the satel­lite would be visible only for the twenty minutes at dawn and dusk that the sun would illuminate it against a dark sky. So the proposal for Project Orbiter included no plans for radio tracking, a decision that—seemingly— was to tell against von Braun less than a year later.

Von Braun sent the proposal to the Department of Defense, JPL, and branches of the armed services. Whipple approached the National Science Foundation and the National Academy of Sciences.

The proposal came at a time when others with influence were strongly advocating that the U. S. should begin a satellite program. Among these were the scientists who had gained the support of the international scientific community for the inclusion of a satellite in the IGY, and, secretly, those in the Air Force who wanted satellites for reconnaissance. In addition, the Naval Research Laboratory (NRL), in Washington D. C., made its own pitch for a satellite program and launcher.

The NRL had been one of the groups to receive V2s, and under the technical leadership of Milton Rosen, it had developed its own rocket—the Viking—for upper-atmosphere research. Rosen persuaded a few forward­looking thinkers to write about the job that satellites could do when rock­ets were developed that could reach orbit. Among these was John Pierce, from AT&T’s Bell Laboratories, who, along with Harold Rosen from the Hughes Aircraft Company, would later be christened “the father of com­munication satellites” by the science fiction writer Arthur C. Clarke.

It was the IGY’s scientists who won Eisenhower’s public support, and on July 29, 1955, the president announced that the U. S. would launch a satel­lite as part of the IGY. He deferred a decision as to which of the armed services would provide the launch vehicle.

The day before the formal announcement, the White House press office summoned journalists, telling them that the briefing was embargoed for the next day. The day’s grace was to allow American scientists time to notify their colleagues overseas. One scientist flew to New York to give a letter for Sydney Chapman, president of the IGY, to a friend who was fly­ing to London.

In a two-hour background briefing at the White House, James Hagerty, Eisenhower’s press secretary, announced that America planned to launch ten basketball-sized satellites sometime during the International Geophysical Year. The journalists rushed to the phones, only to be brought up short by the locked doors. Hagerty was determined that the journalists should understand that the story was not to be broken until after the official announcement the next day. One journalist asked what he, a crime reporter, was supposed to do with the story.

Jet Propulsion, the sedate journal of the American Rocket Society, stopped the presses. The editor wrote, “We cannot imagine anything more exciting than to look up and see through our binoculars the brilliant point of light in the sky that will represent a new astronomical body created by man. We expect to be deeply moved.”

Until the announcement, only about a hundred people knew of the satellite plans. Now the secret was out. But the launch vehicle decision was still pending.

The unenviable task of recommending a launch vehicle fell to a panel of scientists headed by Homer Joe Stewart. Stewart, widely referred to as Homer Joe, was well respected for his technical expertise. He was an aerodynamicist from the California Institute of Technology and a prime mover on the Upper Atmosphere Research Panel. Homer Joe s panel was guided in its deliberations by two precepts: satellites must not interfere with missile development, and the project should have a strong civilian flavor and scientific component. Both charges to the Stewart committee must surely have stemmed from the more secret deliberations within the White House that the Killian report stimulated.

The first requirement was largely taken care of when the Air Force, which had earlier been assigned responsibility for military space projects, dropped out of the competition, saying that it could not develop a rocket in time to launch a satellite during the International Geophysical Year without compromising its missile development. Not a surprising decision, given that the development of intercontinental ballistic missiles had top priority with the Department of Defense and that parts of the Air Force were lobbying for funding to develop reconnaissance satellites.

The second requirement was more difficult to fulfill. Only the mili­tary were then sponsoring launcher development.

But the Army and the Navy squared off: the ring, the committee rooms ofWashington D. C. and private offices in the Pentagon.

In one corner was Wernher von Braun with an updated version of Project Orbiter. At his back, the formidable figure of General John Medaris and the technical expertise of the Jet Propulsion Laboratory. The laboratory had suggested important changes to the upper stages of the launcher, but the size of the proposed satellite—five pounds—stayed the same.

Nevertheless, the Army’s proposal had an immediate strike against it, because the Army was developing intermediate-range ballistic missiles. While not as high a priority as those that the Air Force was developing, they were undoubtedly missiles and compromised perception of Orbiter as a civilian project.

In the opposing corner was the Naval Research Laboratory with its proposal, called Project Vanguard. That proposal contained a much stronger scientific component than the Army’s, a component that the Naval Research Laboratory, with its strong background in upper atmo­sphere research, was well qualified to implement.

The laboratory had additional advantages. It had successfully devel­oped the Viking sounding rocket as a replacement for the V2. The Viking was not part of a weapons system but was intended for scientific explo­ration of the upper atmosphere—albeit in support of military purposes.

Further, the Viking project had provided the lab with a record of developing a large project on time and to cost. The Glenn L. Martin Company had been the contractor for Viking, and the laboratory pro­posed to work with them again on Vanguard.

Finally, the NRL had what was for that time a sophisticated radio tracking plan, whereas the initial Project Orbiter included only optical tracking.

Much of Project Vanguard was technically innovative, relying on miniaturization to beef up the satellite’s capabilities. While miniaturization is today a technological cliche, in 1955 the concept was new. Vanguard would have tiny batteries and radio equipment and in principle would be able to carry ten pounds of scientific instruments, far more than Project Orbiter offered. The batteries were chemical because the designers decided that solar-battery technology, generating electricity from sunlight, was unlikely to be practical in the time available.

The Navy won the first round, but the Army bounced back with a plan that took account of technical criticisms, including an improved tracking proposal. Milton Rosen, who had headed the Viking develop­ment and would be the technical director of Project Vanguard, listened to them argue their case for most of an afternoon. He watched tensely, afraid that von Braun’s eloquence was swaying the panel.

In the event, Homer Joe’s committee confirmed its original decision, and on September 9, the secretaries for the Navy, Army, and Air Force were notified that the Navy would head the three services in the develop­ment of a satellite launcher. When the news reached the NRL’s site by the Potomac in Washington D. C., there was jubilation, some surprise, and a conviction that they could do the job.

The decision of Homer Joe’s panel was not unanimous. Homer Joe himself was one of two who thought that the Army should get the job. The nation could save time and money, he thought, by piggybacking the satellite program on the Army’s missile work. Many, then and now, believe that he was right, and that if the U. S. had picked Project Orbiter, America would have been the first into space.

Thus Project Orbiter became famous as the proposal that could have begun the space age but did not. In retrospect, and in the light of the emerging evidence concerning the Eisenhower administration’s desire to establish the freedom of space by launching a civilian scientific satellite, one suspects that the Army could not have won. Certainly, the guidance given to the Stewart committee seems to have been fashioned to favor the somewhat less militaristic Navy proposal.

The administration’s decision to back Project Vanguard has pulled down much vilification on President Eisenhower’s head. Though it seems now that Eisenhower’s decision was more subtle than it appeared at the time, he did underestimate the blow to America’s self-esteem and the con­comitant gain in the Soviet Union’s international prestige that would fol­low the Sputnik launch.

During 1955, as wheels of policy turned within wheels, the scientists who in the fall of 1954 had won backing from the international scientific com­munity for the inclusion of satellites in the IGY had also been busy. Early in 1955, the executive of the national committee had established a rocketry panel to evaluate the feasibility of producing a rocket that could launch a satellite. The task was delegated to William Pickering, Milton Rosen, and a young hawk, also from the Naval Research Laboratory, John Townsend.

They met in Pasadena in February of 1955. This trio concluded— unsurprisingly—that a satellite launch was feasible. Milton Rosen reported their findings to the rocketry panel in March 1955. He outlined three pos­sible rocket configurations, which, unbeknownst to many scientists of the IGY, had much in common with proposals which were then vying for attention at the Pentagon.

On March 9, Joseph Kaplan, chair of the U. S. IGY, sought approval from the U. S. executive committee for the IGY for the inclusion of satel­lites in their research plans.

He met opposition. The debate, recorded in scribbled, handwritten notes, is hard to decipher, but the language suggests that some of the scien­tists knew or had an inkling that national security issues were at stake. Merle Tuve, director of the Department for Terrestrial Magnetism at the Carnegie Institution (and founder of the Applied Physics Laboratory, which would develop the Transit navigation satellites), was very uncom­fortable with the idea of tying the satellite project to the IGY. Others, like Fred Whipple and Harry Wexler, who as chief sci’entist of the weather bureau would be a strong supporter of meteorology satellites, argued persuasively that satellites would expand the science that could be undertaken by the IGY.

His colleagues asked Tuve—wouldn’t it be better for satellites to be part of the IGY so that the data collected by the spacecraft would be unclassified and freely available? If the Department of Defense were to lead the way, there would be less opportunity for international collaboration. They could do an enormous amount of science with satellites, they argued. And if they were to exclude everything from the IGY that had a possible military application, there would be nothing left. Who were they, asked one scientist, to hold up progress?

Hugh Odishaw, the indefatigable secretary of the national commit­tee, finally summarized their options: say no, irrespective of whether the satellite project had scientific merit or not; embrace the satellite project as part of the American contribution to the IGY; decide whether it had merit and, if it was geophysically useless, say that the satellite program was more appropriate to the Department of Defense.

Clearly, satellites do have considerable potential for geophysics, and what seems to have persuaded Tuve to endorse the satellite project was that if the Pentagon took the initiative, cooperation with other countries would be difficult. Eventually, the executive committee approved the inclusion of satellites in the IGY.

The National Science Foundation then requested funds from Con­gress on behalf of the National Academy of Sciences, which was organizing the IGY (the National Academy of Sciences does not request funds directly from Congress). The National Science Foundation and the National Acad­emy of Sciences gave Donald Quarles, assistant secretary of defense for research and development, what he needed, two bodies with indisputable credentials in the world of civilian science. He now had a formal request for a civilian satellite as had been recommended by the Killian committee.

Kaplan sent a letter to the National Science Foundation explaining that the satellite project would need $10 million in addition to the nearly $20 million that congress had by now authorized for the IGY. At the same time, the IGY s satellite proposal was sent to the White House and the Pentagon. At the Pentagon, Quarles looked at Kaplan’s figures and doubled them. Even that turned out to be a serious underestimation of the ultimate cost (elevenfold what Kaplan requested, excluding the Pentagons expendi­ture), inevitably, perhaps, because no one had any idea of the technical dif­ficulties and costs of space exploration.

It was now May 1955, and the IGY’s organizers needed the money as soon as possible if they were going to launch a satellite before the comple­tion of the IGY at the end of 1958. But the project stalled. At least, that is how it appeared to the staff of the U. S. national committees secretariat, who were frustrated as phone calls went unanswered and unsatisfying memos arrived in response to requests for information about the satellite budget.

In fact the IGY was now a pawn in a bigger game. The chain of events was in progress that led ultimately to the National Security Coun­cil’s endorsement of the plan at the end of May, to the formation of Homer Joe Stewart’s committee, to President Eisenhower’s announcement of July 29, and to the September decision that Project Vanguard would launch America’s first satellite.

From the IGY’s perspective, things started to move again after Eisen­hower made his announcement. Despite not knowing which launch vehi­cle the administration would select, the national committee asked Richard Porter, a technical consultant to General Electric, to form a panel to over­see the IGY’s satellite work. Porter’s committee spawned two more. One, led by William Pickering, was responsible for radio and optical tracking and orbital computations. Fred Whipple was one of his committee mem­bers, and he took on the responsibility of organizing optical tracking for the IGY. James Van Allen chaired the second group, which was responsible for choosing the experiments that would fly.

As 1956 advanced, the cost of the satellite program escalated. Eventu­ally, one of the IGY scientists complained that the satellite program was costing as much as if each satellite had been built from solid gold. Berkner assured him that the value of the satellites to the international community and to the United States was worth far more than digging that much gold out of the earth.

Despite such enthusiasm, development of the Vanguard rocket was experiencing technical and financial difficulties. Milton Rosen, Van­guard’s technical director, was convinced that the Glenn L. Martin Com­pany was not giving the NRL its best efforts. He attributed this to the company’s also having the contract for the Air Force’s Titan ICBM. Whatever the reasons, the pitfalls and colossal nature of undertaking to develop and build a three-stage rocket from scratch in so short a time were now apparent.

These technical difficulties were compounded by the cutbacks in the program. By October 1957, Van Allen’s committee could count on only six launch vehicles, possibly far fewer.

Sputnik was about to change all that.

On Friday, October 4, 1957, Walter Sullivan, a science correspondent for the New York Times, filed a story for Saturday’s paper. After a week at the rockets and satellite conference he felt fairly certain that he had taken the right line. He’d written that the Soviets were close to a launch attempt.

Sullivan says that before going to the reception he called his office. He was told that the paper’s Moscow office had heard Radio Moscow announce that a satellite was in orbit. He was told to find out what he could.

Sullivan hurried to the embassy. He found the American scientists: Pickering, Richard Porter, Lloyd Berkner, and John Townsend. When Sul­livan told them the news, they looked at one another in consternation. To this day, Pickering believes that the Soviets at the reception did not know of their own successful launch.

The Americans went into a huddle and decided that they must con­gratulate their hosts. As the senior American scientist present, the task fell to Lloyd Berkner. Leaving his colleagues to assimilate the news as best they could, he took his glass and a spoon and climbed on a chair. He rapped the glass for attention. Slowly the conversation died down, and the guests turned to Berkner. His announcement was simple, and it caused a sensa­tion. He said, “I am informed by the New York Times that a satellite has been launched and is in orbit at an altitude of nine hundred kilometers. I wish to congratulate our Soviet colleagues on their achievement.” With that, he raised his glass. Blagonravov was beaming as he drank the toast.