Category Something New Under the Sun

A Time of Turbulence

This dread and darkness of the mind cannot be dispelled by sunbeams, the shining shafts of day, but only by an understand­ing of the outward form and inner working of nature…

First, then, the reason why the blue expanses of heaven are shaken by thunder…

As for lightning, it is caused when many seeds of fire have been squeezed out…

The formation of clouds is due to the sudden coalescence…

—Lucretius, On the Nature of Things

L

ucretius sought rational, deterministic explanations for the weather.

These turned out to be wrong, but one suspects that the Roman philosopher may have guessed this for himself. He wrote that it was better to venture on an incorrect rational explanation than to submit to supersti­tion: no sacrifices for him to propitiate the gods. And no sacrifices, except of time and effort, for those who during the past hundred years or so have wrestled to turn meteorology into a science.

For most people—farmers, sailors, or those of us going about our ordinary business—meteorology means and has always meant the weather forecast: the difference between heading for the golf course or curling up at home with a good book, between planting crops or waiting, and ulti­mately for some the difference between life and death. Those forecasts, dispensed in a few minutes on nightly news broadcasts, rest on the integra­tion of a staggering amount of mathematics, physics, engineering, and computer science. In the first century B. C.E., while incorporating his own ideas with the philosophy of Epicurus and turning the whole into verse, Lucretius was at a considerable disadvantage.

Only in the nineteenth century did the modern era of weather fore­casting begin. The introduction of the telegraph allowed observers to communicate to those at distance points what weather was coming their way. Such timely reporting also allowed meteorologists to plot weather
maps and to develop the concept of storm fronts and cyclones. From the 1920s, radio balloons collected readings of temperature, wind speed, pres­sure, and moisture content, improving knowledge of conditions at altitudes in the lower atmosphere. Later, in the 1950s and 1960s, scientists took the important step of incorporating knowledge of the upper atmosphere into their understanding of meteorological conditions in the lower atmosphere, that is, they explored how the upper atmosphere affects weather at the sur­face.

But until the middle of the twentieth century, meteorology was only slowly breaking free of its ancient reliance on folklore and supersti­tion. It was still more of an art than a science. Then came computers, mathematical modeling of atmospheric behavior, and weather predictions based on computer models. Gradually, it became possible to combine and manipulate observations from many different sources—from ocean buoys to Doppler radar and satellites.

Weather satellites inserted themselves into this history as best they could—not always felicitously. They were a technology in which some in the 1950s intuitively saw promise because of the unique bird’s-eye view from space, but it was only in the early 1980s that the advocates of satellite meteorology succeeded in winning widespread acceptance from the mete­orological community.

In the very earliest days of satellite meteorology, a few names stand out in what was a tiny, intertwined community. The first are William Kel­logg and Stanley Greenfield, who in 1951 while at the RAND Corpora­tion (consultants to the Air Force) published the first feasibility study on weather satellites. Then came Bill Stroud and Verner Suomi, who com­peted to have their experiments launched on one of the satellites of the International Geophysical Year. Each, after vicissitudes, flew an experi­ment. Stroud’s failed, because the Vanguard satellite that carried it into space was precessmg wildly. Stroud went on to head NASA’s early meteo­rological work at the Goddard Space Flight Center and to argue the case for satellite meteorology at congressional hearings. Suomi’s satellite pro­duced data, and he remained in the trenches of science and engineering, making frequent forthright forays into the policy world both nationally and internationally.

There were also Harry Wexler and Sig Fritz from what was then called the Weather Bureau. Wexler, who died in the 1960s, is someone whose name in this context is often forgotten, but as chief scientist of the

Weather Bureau and an active participant in the committees planning the IGY, he was an important supporter of satellite meteorology. He was one of the scientists arguing persuasively in the face of Merle Tuve’s doubts that the IGY should include a satellite program. And Wexler was a staunch ally of a belated attempt by Verner Suomi to participate in the IGY, drum­ming up support for Suomi from eminent meteorologists like Kellogg at RAND.

Fritz worked for Wexler. When the Weather Bureau set up a satellite service, Fritz was its first employee. He was assigned office space in a cleaned out broom cupboard. There, undaunted by the Vanguard failures and the modesty of his office space, Fritz worked with NASA on the first American weather satellite—TIROS. Both Wexler and Fritz were consul­tants for Verner Suomi’s IGY experiment.

Fritz recruited Dave Johnson,[9] who, like Suomi, became an out­spoken proponent of satellite meteorology. Johnson eventually headed the satellite division of what, after several bureaucratic incarnations, was to become the National Oceanic and Atmospheric Administration.

Except for Kellogg and Greenfield, these men worked in the civilian world but also made forays into the “black” world of defense projects, namely the Air Force s Defense Meteorological Satellite Program. The Air Force was an important player in the history of satellite meteorology, devel­oping both engineering and analytical methods for interpreting satellite imagery. And the participation of people like Johnson in both worlds pro­vided a conduit, albeit of limited capacity, for technology transfer from mil­itary to civilian satellites. The story of this important part of the history of satellite meteorology—the way that the defense and civilian worlds inter­mixed—will have to wait until all the relevant documents are declassified.

Despite the limitations imposed by not having a full understanding of the interplay between civilian and defense projects, some broad aspects of the history of satellite meteorology are clear. It is a more complicated story than that of satellite navigation, mainly because it is the story of a technology being developed for a field that was still transforming itself from art to science.

One of the most outspoken and energetic participants in the field’s history was Verner Suomi, of the University of Wisconsin in Madison. Some have called him the father of satellite meteorology.

In 1992, Dave Johnson, then working for the National Research Council of the National Academy of Sciences, recalled a meeting of the world’s leading meteorologists in 1967 when they were planning an inter­national effort, known as the Global Atmosphere Research Program, to study the atmosphere. GARP eventually got underway in the late 1970s. Suomi’s task was to summarize the specifications that weather satellites would have to meet in order to fulfill GARP’s research goals. Johnson said: “We threatened to lock Vern in a room and not to let him have food or drink until he’d written everything down. We didn’t, of course, but he hated writing, and we had to keep an eye on him.”

Suomi’s colleagues were wise to put pressure on him. During late 1963 and early 1964, when Suomi spent a year in Washington DC. as chief scientist of the Weather Bureau, he claims to have written only four memos—which may be the all-time minimalist record for a bureaucrat.

One of GARP’s roles was to set research priorities given what were then the comparatively new technologies of high-speed computing, math­ematical modeling, and satellites. Those priorities give a sense of the immensity of the task facing meteorologists.

The priorities were:

• Atmospheric composition and structure;

• Solar and other external influences on the earth’s atmosphere;

• Interaction between the upper and lower atmosphere;

• Interaction between the earth’s surface and the atmosphere;

• General circulation and budgets of energy, momentum, and water vapor;

• Cloud and precipitation physics;

• Atmospheric pollution;

• Weather prediction;

• Modification of weather and climate (no longer popular);

• Research in sensors and measuring techniques.

A study of these topics would need the “observation heaped on observation” that Sir Oliver Lodge spoke of in his lecture about Johannes Kepler: some observations were to be made by radar, others by airline pilots, weather balloons, and ground-based instruments. And some, of course, would be recorded by satellites.

Despite the vibrancy of meteorological research typified by plans for GARP, it was clear by 1967 that persuading the wider meteorological community—both line forecasters and many research meteorologists—to accept data from satellites would be an arduous task.

Many of the important steps to acceptance were choreographed, in part at least, by Suomi or Johnson and the groups that they headed. Nei­ther man was shy in his advocacy of the technology. Johnson, in fact, threatened on one occasion to “blow his stack” with his boss, whom John­son felt was hostile to satellite data. None of the advocates of satellites could afford too many niceties. The money spent on weather satellites prompted resentment from many. And there were reservations and criti­cisms about satellite meteorology.

Part of the opposition lay, as always, in suspicion of a new technology. But part of it was due to the technology’s acknowledged limitations, which were (and are) imposed by the nature of satellite observations. Satel­lites do not directly measure the meteorological parameters—tempera­tures, pressures, wind speeds and moisture contents at as many latitudes, longitudes, and altitudes as possible—that are essential for computer mod­els and any quantitative predictive understanding of the atmosphere’s behavior. Instead, satellites “see” visible and infrared radiation welling up from the earth. Meteorologists thus have either images or radiometric measurements as their raw data, and from these they must infer quantita­tive meteorological parameters. The inferences are not easy to draw. They call for considerable knowledge of atmospheric physics and chemistry and rely on clever mathematical manipulations of the equations describing atmospheric behavior.

Images rather that radiometric measurements came first in the his­tory of meteorology satellites. Kellogg’s and Greenfield’s study of satellites for “weather reconnaissance,” which was carried out before numerical weather prediction had become central to the future of weather forecast­ing, envisaged that spacecraft would carry still cameras aloft. These would photograph cloud cover, and meteorologists would then study the cloud types and distribution in a qualitative attempt to gain insight into atmo­spheric behavior and thus improve weather forecasting. In the course of their study, Kellogg and Greenfield posed some of the important questions that would preoccupy early satellite meteorologists. These were:

• How could you tell which bit of the earth the camera was looking at and thus where the cloud cover was?

• How could you tell what type of clouds you were looking at and what their altitudes and thicknesses were, and thus what signifi­cance they had to a developing weather system?

• How could you get the information to line forecasters in a timely fashion? It would not be much use telling a ship that there had been an eighty percent chance of a storm yesterday The launch of the first weather satellite—TIROS I (for thermal infrared and observing system)—in April 1960 confirmed that these were all tough and legitimate concerns.

Nevertheless, TIROS showed for the first time what global weather patterns looked like. The promise inherent in the technology was there for all to see in grainy black and white. But it convinced only those who already believed. Succeeding satellites in the TIROS and improved TIROS series carried gradually more sophisticated instruments, each of which slowly took satellite meteorology closer to wide acceptance.

One such class of instruments—known as sounders—were first developed by Johnson’s group in the 1960s. Sounders measure temperature and, more recently, the moisture content of the atmosphere at different altitudes and in places where direct measurements with, say, a thermometer are not possible—over oceans, for example, where much of the weather develops. They are important for near-term predictions of severe weather such as thunderstorms.

The sounder relies on inferences made from radiometric readings at different frequency ranges in the infrared portion of the spectrum and on its operators’ detailed knowledge of atmospheric chemistry and physics. Inevitably, there is greater inaccuracy in the values of temperature and moisture content taken from satellite sounders than from direct measure­ments of the same parameters. And so modelers have, for the most part, not liked to rely on data from satellite sounders. A notable exception is the European Center for Medium Range Weather Forecasting, which has taken the lead in finding ways to extract from satellites the information that is needed for computer models. By the early 1990s, the center was saying that satellite soundings had extended useful predictions from five and a half to seven days in the Northern Hemisphere and from three and a half to five days in the Southern Hemisphere.

While Johnson’s group developed the first sounder, Suomi came up with the idea for the spin-scan camera, which flew for the first time in 1966. Although this class of camera was to become a crucial meteorologi­cal instrument, Suomi was told by a colleague ten years after it first flew that if submitted as part of a Ph. D. thesis, it would not merit a doctorate.

Thus satellites were not entirely welcome participants in meteorol­ogy. Far more welcome were the new high-speed computers and John Von Neumann’s conviction that with sufficient computational power one could model the atmosphere’s behavior and predict the weather.

The idea for such numerical weather prediction was proposed first in 1922 by Lewis Richardson. He tested his idea by feeding meteorological data that had been collected at the beginning of International Balloon Day in May 1910 into mathematical models describing atmospheric behavior. He compared his numerical predictions with the data collected during the day and found no agreement. Discouraged, Richardson concluded that to predict the weather numerically one would need 64,000 mathematicians who would not be able to predict weather conditions for more than sec­onds ahead; they would, in effect, be “calculating the weather as it hap­pened.”

In the thirty years following Richardson’s depressing experience, much changed, including improved understanding of the physics of the atmosphere and mathematical analysis of its behavior. Thus, when the tech­nology of computing emerged, modelers set to work, weaving the basic physical laws into models mimicking the behavior of the atmosphere. And the computers took over the calculations. Initially, the models represented only surface events in small regions. Subsequently, modelers incorporated the influence of the upper atmosphere on weather at the surface.

There are now many models—global, hemispheric, regional. Some are mathematical behemoths constructed from thousands of equations. Some give short-term weather predictions, while others look up to two weeks ahead—so-called medium-term forecasts. Yet others make forecasts, extremely controversial ones, far into the future as climatologists explore climatic change.

All, however, devour numbers—values of temperature, pressure, etc. And because the early satellites did not supply the quantitative data that the models required, there was tension between computer modelers and satellite advocates. Both groups, after all, were seeking scarce public funds for expensive technologies.

In 1969, ten years alter the first meteorological payload was launched, the National Academy of Sciences wrote, .. numerical weather prediction techniques demand quantitative inputs, and until weather satellites are able to generate these, their use in modern meteorol­ogy will be at best supplementary.”

Nearly thirty years later, the technologies have become more com­patible and weather satellites have obtained a secure place in meteorology. The Air Force, NOAA, NASA, and academic groups like that of Suomi’s at Wisconsin have done what they can to extract meteorological values from unprepossessing streams of satellite data and, importantly, to make this information compatible with observations from weather balloons, radar and surface instruments. Yet, says Johnson, considerably more information could be extracted from the meteorological satellite data.

Weather satellites gather their data—images and soundings—from two different types of orbit: polar and geostationary. Like Transit, a weather satellite in polar orbit follows a path that takes it over the poles on each orbit, while the earth turns through a certain number of degrees of longi­tude m the time it takes the satellite to complete one orbit. Thus polar – orbiting satellites, if they have a wide enough field of view to either side of the subsatellite point, provide global coverage. Their altitude, and thus how long they take to complete an orbit, is chosen so that the satellite will “see” all parts of the earth once every twelve hours.

To be truly useful, however, weather satellites need to occupy a special kind of polar orbit, known as sun-synchronous. Sun-synchronous orbits are chosen so that the satellite maintains the same angular relationship to the sun, which means that the satellite will be above the same subsatellite point at a given time of day. Its readings are then consistent from day to day. The timing of the orbit is chosen so that the satellite readings are available for the computer prediction models, which are run twice a day.

If the orbit is to maintain the same angular relationship to the sun throughout the year, it cannot remain fixed in space. But orbits are not, of course, fixed. They respond to the earth’s gravitational anomalies. Mission planners achieve sun-synchronous orbits by exploiting the known effects of the earth’s gravitational field. They select inclinations and altitudes that result in the orbit moving in such a way that the satellite’s sun-synchronous position is maintained. The consequences of the natural world that the Transit team had to understand and to compensate for can thus be exploited usefully by those planning the orbits of weather satellites.

The laws of physics result, too, in the existence of the extremely use­ful geostationary orbit. A satellite at an altitude of about 36,000 kilometers takes twenty-four-hours to complete an orbit. If the orbit has an inclina­tion of zero degrees, that is, the plane of its orbit is coincidental (more or less) with the plane of the equator, then the satellite remains above the same spot on the earth. Thus, the satellite is with respect to the earth for all practical purposes stationary and can view the same third of the earth’s surface while the weather moves underneath it. Suomi’s spin-scan cameras were designed for this orbit. Geostationary orbits were also to prove of critical importance to communication satellites, and Suomi’s spin-scan camera was first launched aboard a satellite designed by one of the fathers of communication satellites—Harold Rosen.

While they are crucial to the beginnings of satellite meteorology, the issues mentioned so far scarcely scratch the surface of the history of weather satellites. There was also an important battle in the early 1960s between NASA and NOAA’s forerunner about the technology of the satellites to replace TIROS and about who would pay for operational satellites. Finally, an improved version of TIROS was selected, and NASA developed the alternative proposal, a more experimental satellite series dubbed Nimbus.

White recalls, “On the same day I was sworn in as chief of the Weather Bureau, Herbert Holloran, the assistant secretary for science and technology, took me to one side and said we have to make a decision about Nimbus. The issue was would we be willing to use Nimbus as our operational satellite. The cost would have been two to three times the cost of using TIROS. This was important to the weather satellite program. If we had followed Nimbus, the cost would have skyrocketed, and maybe we wouldn’t have got the money from Congress. We decided on the basis of cost to go with TIROS. I think that was the right step.”

Even from a technical standpoint, the history is not straightforward. There was no single event, such as Guier and Weiffenbach’s tuning into Sputnik’s signal, from which the story unfolds. Nor was there one clearly defined technical goal such as that of the Transit program—locate position with a CEP of one tenth of a mile. All of the physics and engineering that went into the Transit program were harnessed to meet that goal and were refined to enable the subsequent improvements in the system. In the field of meteorology, satellites were just one tool wielded to learn more about the atmosphere, and no one really knew what needed to be learned as is apparent from the breadth of The Global Atmosphere Research Program’s aims. It is, therefore, not surprising that meteorology satellites took longer than navigation spacecraft to find acceptance.

In further contrast to Transit, there was no single group, like the Navy’s Special Projects Office, that wanted weather satellite technology. Even the Weather Bureau, outside of Johnson’s group, was unenthusiastic. Further, no single group, like the Applied Physics Laboratory’s Transit team, was central to the development of weather satellites. True, the Air Force, backed by sundry laboratories and consultants such as the RAND Corporation, was interested from early days, but once the IGY’s satellite program was announced, more scientists became involved, including Verner Suomi and Bill Stroud. After the launch of Sputnik I, the Advanced Research Projects Agency sponsored the TIROS program, which NASA took over when that agency opened its doors in October 1958. Industry, including companies like RCA, took a hand, and, of course so did the Weather Bureau.

If the professionals were slow to accept meteorology satellites, the lay audience was intrigued by the potential of a spacecraft’s global view, and popular articles appeared in the newspapers of the 1950s speculating on the importance of satellites for weather forecasting. They pointed out that only satellites would be able to provide comprehensive and frequent read­ings over the approximately seventy-five percent of the earth’s surface that is covered by ocean.

Since the first TIROS went into orbit, the United States has launched more than one hundred meteorology satellites. Now the coun­tries of the former Soviet Union, Europe, Japan, the People’s Republic of China, and India maintain meteorology satellites. All contribute to the global economy by improving forecasts for agriculture and transport, and to safety by monitoring severe weather such as hurricanes and allowing more timely and accurate predictions of where they will make landfall. It is unlikely, in the U. S. at least, that a hurricane will ever kill more than 6000 people, as did the hurricane that struck Galveston, Texas, in 1906. It has taken more than three decades, but weather satellites are now living up to the popular expectations of the 1950s.

Chapter five: Polaris and Transit

Information about Polaris (pp. 49 and 51-53), its purpose and develop­ment, emerged during many hours of interviews with members of the Transit team.

Books that provided background for chapter five include The Polaris Sys­tem Development: Bureaucratic and Programmatic Success in Government, by Harvey M. Sapolsky (Harvard University Press, 1972); Forged in War: The Naval—Industrial Complex and American Submarine Construction, 1940—1961, by Gary Weir (The Naval Historical Center, 1993).

The potted history of navigation in this chapter (pages 50 and 51) draws on interviews with Commander William Craft and Group Captain David

Broughton, and on From Sails to Satellites:The Origin and Development of Navigational Science, by J. E.D. Williams (Oxford University Press, 1992).

Transit’s status as brickbat-01 and the meaning of this terminology (page 53) emerged during interviews with Transit team members.

The fact that radars capable of detecting a periscope’s wake were being developed at the end of the 1950s and in the early 1960s (page 54) comes from George Weiffenbach.

How the technology of the Transit receivers evolved (page 55) comes from Tom Stansill and from papers and old sale brochures that he sent to me.

The history of APL (page 56) comes from The First 40 Years, JHU APL (Johns Hopkins University Press, 1983).

Notes about Frank McClure’s and Richard Kershner’s previous careers (page 56) come from The First 40 Years, JHU APL and from press releases and briefing papers sent to me by Helen Worth, the APL’s press officer.

The fact that Ralph Gibson was sounded out as a potential first director of the Defense Research Establishment is mentioned by Admiral William Raborn, head of the Fleet Ballistic Missile program until 1962, in an oral history at the Naval Historical Center, in Washington, DC (page 57).

Chapter twenty: Syncom

Chapter 20 is a distillation of information from the following documents, presented in chronological order; though extracts from one document or an appendix may have been useful for explaining some other part of the unfolding events.

Interviews with Rosen and Roney provided detail that do not appear in the written record. I have included it where it seemed to make sense. For example, it was Rosen who told me that T. Keith Glennan first told Puckett that he was “talking through his hat” when Puckett presented the company’s idea for a 24-hour satellite (page 212). Though Glennan was interested in what Puckett said, Glennan knew that НАС then knew nothing about satellites and that the 24-hour satellite idea was far from conservative. Glennan might well have said what Rosen recalls he said.

April 26, 1960 Rosen completes evaluation of life and reliability of the proposed satellite, focusing on electronics and TWT design. A handwrit­ten note from Puckett says, “This looks very promising. Thanks.”

May 1960: A hefty, mathematical document from Williams entitled “Dynamic Analysis and Design of the Synchronous Communication Satellite.”

A memo from J. W. Ludwig to C. G. Murphy and A. E. Puckett of May 2, 1960, discusses a meeting with E. G. Witting, of the Army, and a represen­tative of the Office of Defense Research and Engineering (Mr. Evans). Hughes learned that the Army was already considering other 24-hour satellite proposals and that Herb York, DDR&E, was “intensely interested in the Hughes program.”

On May 19,J. W. Ludwig sent a memo to A. E. Puckett about a forth­coming request for proposals (July 1, 1960) from the Army for the Advent 24-hour communication satellite.

June 1960: Synchronous communication satellite, proposed NASA exper­imental program by the НАС Airborne Systems Group. Proposal included details of Jarvis Island where НАС was at that time proposing it should build a lunch site.

A memo from Lutz for file copied to Rosen on June 3, 1960, summarizes in detail presentations by Pierce, Jakes, and Tillotson from Bell Telephone Laboratories at a conference at the end of May.

Letter from Douglas Lord, technical assistant to the Space Science Panel, thanking Allen Puckett for the Hughes presentation to the President s Science Advisory Committee.

2 77

Harold Rosen told me the anecdote of how the meeting came to be set up (page 212).

July 26, 1960, Puckett confirms a meeting requested by Abe Silverstein for Hughes to present its satellite proposal to Keith Glennan.

Technical memos in the meantime show Williams’s preoccupation with dynamics studies and his involvement with NASA’s Langley field center.

Memo from John Richardson to C. G. Murphy of August 12, 1960, talks of GT&E’s evaluation of the Hughes satellite. Richardson believed that the GT&E reaction was quite good. GT&E asked Hughes for justification of the life expected for the TWT; how the telemetry system could be protected from disruptive tampering and how the Hughes assertion that ten launches would be needed for one successful satellite could be recon­ciled with the company’s plans for three launches.

An internal memo from Ralph B. Reade to Roy Wendahl raised concern about internal conflicts in the field of satellite communication and the company’s fragmented approach to military customers.

September 14, 1960: preliminary cost estimates of a commercial commu­nication satellite. Total: $15.75 million, including $4 million for the Jarvis Island launch site.

A letter from John Rubel to A. S. Jerrems dated September 22, 1960, refers to a visit he had made a few weeks earlier to Hughes.

On October 25, 1960, Rosen wrote back to Witting refuting all the spe­cific criticisms that Witting made of the НАС proposals. Rosen con­cluded, “In our opinion, the Hughes proposal, if implemented, would achieve all the objectives of the present program, but at an earlier date and lower cost.”

Memos in September, October, and November show that НАС held meetings with GT&E, The Rand Corporation, Bell Telephone Laborato­ries (November 2, 1960), ITT, and British Telecommunications.

A Letter from Allen Puckett to Lee DuBridges, president of Caltech, written on November 18, 1960, refers to their discussions about commu­nication satellites at the Cosmos Club.

Memo from Allen Puckett to J. W. Ludwig of December 14, 1960, con­cerning NASA’s forthcoming RFP for a medium-altitude active satellite.

Memo from Bob Roney to Allen Puckett on December 23, 1960, con­cerning the need for budget decisions for the 24-hour communication satellite program for 1961.

In a memo of January 6, 1961, Puckett informed Rosen, Williams, Hud­speth, and a few others that Lutz would once again be evaluating their communication satellite proposal.

An agenda of an Institute of Defense Analysis meeting shows Rosen scheduled to brief the group between 10:30 and 11:00 on January 10, 1961. Rosens write-up of the meeting on January 12 was directed at Allen Puckett. I noted that the panel was critical of Advent and that his description of the Hughes 24-hour satellite elicited “generally favorable comment.”

In a memo to C. G. Murphy and F. P. Adler, dated January 13, 1961, Sam Lutz demonstrated a less enthusiastic view of the 24-hour satellite, advo­cating extensive additional work and comparisons with passive and medium-altitude active satellites.

Memo from S. G. Lutz to Allen Puckett and A. V. Haeff of February 10, 1961. Subject: review of satellite communications. Lutz wrote a negative and critical report of the 24-hour satellite.

Telegram from Rosen to Rubel of February 28, 1961, apparently responding to remarks by Bill Baker, of Bell Telephone Laboratories.

March 1961: НАС report, “Stationary Satellite, Island Operation Phase.”

Memo from Allen Puckett to F. P. Adler of April 20, 1961. Subject: a con­versation Puckett had had with John Rubel and Rubel’s suggestions regarding the actions that Hughes should take (НАС archives: Commer­cial Communication Satellites 1990-05, Box 6 DoD Communication Satellite).

Memo from C. Gordon Murphy to J. W. Ludwig of April 27, 1961. Sub­ject: A conversation Murphy had had with Rubel.

Letter dated May 8, 1961, from Allen Puckett to John Rubel. Subject: A Program for Interim Satellite Communication.

Memo dated May 11, 1961, from Williams to Hyland, outlining the mis­takes he thought the company had made with respect to developing the 24-hour communication satellite.

Telegram dated May 18, 1961, from Jack Philips to R. E. Wendahl, telling him that Hughes had just been informed that they were unsuccessful in their bid for Project Relay.

Memorandum for the Associate Administrator from Robert Nunn and Leonard Jaffe to Robert Seamans, dated June 6, 1961. The memo states, “It is recommended that NASA immediately embark on a project leading to tests of a simple, light weight communication satellite at 24-hour orbital altitudes. …” This memo formalized a situation that Seamans and Rubel had worked to bring about.

Memo from A. S. Jerrems to Allen Puckett dated June 19, 1961.Jerrems wrote, “Here is another bulletin on the Rubel situation. In a telephone conversation this weekend, Rubel advised me that he spent all day Satur­day (June 17) in a meeting with NASA to discuss communication satel­lite plans. He was inscrutable about the detailed content of the meeting, but he made a statement to the effect that in his opinion, НАС prospects for getting a synchronous satellite funded were better now than they had ever been.”

A letter of June 23, 1961, from Roswell Gilpatric, deputy secretary of defense, to James Webb. He writes, “Mr. Rubel has told me about the plans that he and Dr. Seamans are formulating for an interim synchronous satellite communication experiment with potential for an interim opera­tional capability within the next 12 to 24 months. … I regard the pro­posed program as complementary to those [Rely and rebound (Echo-type satellite)] and the Advent project. … You have my assurance of support in the event that, in your judgement, their proposals should be adopted.” With this letter, Gilpatric set aside the informal agreement that NASA would not develop synchronous-altitude satellites.

Memo from A. S. Jerrems to J. H. Richardson, Allen Puckett, and R. E. Wendahl of July 20, 1961, on the possibility of an early synchronous-orbit experiment. НАС still did not know that both NASA and the DoD were proposing a sole-source contract. After dinner and talks with Rubel, Jer – rems wrote, “Although the proposed five week study program for Hughes, which Rubel and Seamans described to us at the end of June, has not yet been kicked off as we hoped it would be, the planning for the Special Program is not quiescent. There have been a continuing series of meetings between NASA and the DoD to iron out the definition of the ground roles for НАС ”

On August 9, 1961, Alton Jones, project manager at the Goddard Space Flight Center and James McNaul, acting project manager for the U. S. Army Advent Management Agency, signed a contract to be jointly pur­sued by NASA and the DoD for the preliminary project development plan for a lightweight, spin-stabilized communication satellite.

On August 12, 1961, Maj. Gen. G. W. Power, director of developments in the Office of the Chief of Research and Development, wrote to Rosen rejecting his ideas for a lightweight, spin-stabilized communication satel­lite.

Bell Systems was well aware of the promise of communication satellites at synchronous altitudes and also aware of the station-keeping difficulties. In a paper written in November 1962, K. G. McKay wrote an internal paper on the pros and cons of synchronous satellites. He wrote, “The synchro­nous satellite must be placed at a specific point in space with exactly the right velocity and kept there for the life of the satellite. It is a bold con­cept and I am confident that some day it will be achieved” (box 840902 – AT&T archives).

Rejections of the Hughes Proposal

From Maj. Gen. Marcus Cooper, Air Research and Development Com­mand, to Allen Puckett on 3 January 1961: “Since your 3 November visit here… analyzed in detail the Hughes Aircraft Company proposal for a synchronous altitude active communication satellite. In general their find­ings reveal that the proposal is technically marginal in several respects, and tends to be overly optimistic. . . .” Cooper goes on to say that in view of the results expected from Advent, he did not think that “we should pro­ceed further with the Hughes proposal at this time.”

From E. G. Witting, deputy director of research and development, to Mr. J. Bartz, assistant manager, for contracts, writing on October 11, 1960, in response to a letter from НАС of April 22, 1960. Witting writes, “Your proposal has been thoroughly evaluated by the Army and it has been determined that the project would not meet the present requirements of the Army for intercontinental communications.”

Documents that contributed to the general framework of the chapter:

“Syncom (Interim Communication Satellite) Chronology,” possibly pre­pared in the spring or summer of 1963, giving dates of John Rubels involvement with Syncom between 10 April 1961 and 6 February 1963.

Advent chronology from 14 April 1960 to December 1961 (John Rubels papers).

Policy statement for exploitation of НАС communication satellite, undated and unsigned.

Preliminary history of the Origins of Syncom, by Edward W. Morse (NASA Historical Note No. 44) September 1, 1964. Some aspects of this report, about agreements between NASA and the DoD for instance, con­firm details in earlier chapters in this section (John Rubel’s papers).

Although John Rubel was interested in Syncom, others in the Office of Defense Research and Engineering were less keen. Dr. Eugene Fubmi, for example, was not (Author’s interview with John Rubel). In a memo dated March 26, 1962, Fubini was still arguing strongly in favor of Advent (John Rubel’s papers).

R. H. Edwards to D. D. Williams, 19 January 62, “Separation of syncom payload from the third stage” (НАС archives 1987-44 box 1).

“Torques and Attitude Sensing in Spin-Stabilized Synchronous Satellites,” by D. D. Williams, American Astronautical Symposium, Goddard Memor­ial Symposium, March 16-17, 1962.

Post Syncom decision:

Interest in Syncom grew once it had become an official project. An inter­nal НАС memo date 9 May 1962 from C. Gordon Murphy to R. E. Wendahl discussed a visit by the commanding general of the U. S. Army Advent Management Agency, who was interested in HAC’s ability to pro­vide a replacement for the Advent Spacecraft.

By June 18, Robert Seamans was writing to John Rubel about NASA’s plans for a follow-on Syncom program—a five-hundred-pound spacecraft that would permit the “incorporation of 4 independent wide­band transponders, redundant control systems and sufficient on-board auxiliary power to operate the system continuously.” Such a satellite, as a memo from Robert S. McNamara, dated May 23, 1962, shows, would provide a suitable alternative to Advent (John Rubel’s papers).

Memo from John Rubel, deputy DDR&E, to the assistant secretary of the Army, January 25, 1962. Subject: DoD support of NASA—Syncom com­munication satellite test (John Rubel’s papers).

Documents for general background to the communication section

“Telephones, People and Machines,” by J. R. Pierce, Atlantic Monthly; December 1957.

“Transoceanic Communication by Means of Satellite,” by J. R. Pierce and R. Kompfner, Proceedings of the IRE, March 1959 (David Whalen, from George Washington University).

“Satellites for World Communication,” report of the Committee on Sci­ence and Astronautics, U. S. House of Representatives, May 7, 1959.

Project Summary: Project Courier Delayed Repeater Communications Satellite. November 17, 1960 (John Rubel’s papers).

Memorandum for the President, presented to Cabinet December 20,

1960 (David Whalen, from George Washington University).

A Chronology of Missile and Astronautic Events, Report of the Commit­tee on Science and Astronautics, U. S. House of representatives, March 8, 1961.

Special Message to the Congress on Urgent National Needs, delivered to a joint session of Congress, May 25, 1961 (David Whalen from George Washington University).

“Hazards of Communication Satellites,” by J. R. Pierce, The Bulletin of the Atomic Scientist, May/June 1961.

“The systematic development of satellite communication systems,” by K. G. McKay for presentation to the American Rocket Society, October

1961 (box 840902, AT&T archives).

“The Commercial Uses of Communications Satellites,” by Leland S. Johnson (The RAND Corporation, June 1962).

“Aeronautical and Astronautical Events of 1961,” report of NASA to the Committee of Science and Aeronautics, U. S. House of Representatives, June 7, 1962.

“Communication by Satellite,” by Leonard Jaffe, International Science and Technology, August 1962.

“The dawn of satellite communication: a cooperative achievement of technology and public policy,” by John A. Johnson (НАС archives).

“Communication satellites,’"Journal of Spacecraft and Rockets 14 (7),July 1977 pp. 385-394.

“Benefits in Space for Developing Countries,” by Theo Pirard, Aerospace International May/June 1980.

“Satellite links get down to business,” High Technology Magazine, June 1980.

“Rocky Road to Communication Satellites,” draft of material prepared by Barry Miller for a lecture in the early 1980s (НАС archives).

“The History and Future of Commercial Satellite Communication,” by Wilbur L. Pritchard, IEEE Communications Magazine 22 (5), May 1984.

“The Bell System,” Encyclopedia of Telecommunications (Marcel Dekker, 1991).

“The American Telephone and Telegraph Company (AT&T),” Encyclope­dia of Telecommunications (Marcel Dekker, 1991).

“History of Engineering and Science in the Bell System,” in Transmission Technology; edited by E. F. O’Neill (AT&T archives 85-70382).

“The Development and Commercialization of Communication Satellite Technology by the United States,” by George Hazelrigg Jr. Draft in the NASA History Office.

How the World was One, by Arthur C. Clarke (Bantam Books, 1992).

[1] The Eisenhower administration wanted to establish the freedom of space by launching a civilian scientific satellite. Such a satellite would pave the way for America to launch reconnaissance satellites that could fly over foreign territory without eliciting inter­national protest or retaliation. Only a few of the scientists participating in the IGY knew of the administration’s secret purpose, and it is not yet clear who these were. In fact, it is only now that historians are beginning to fully uncover the political relationship between the IGY and the reconnaissance satellite program. See. . . the Heavens and the Earth: a political his­tory of the space age, by Walter A. McDougall (Basic Books, 1985), which contains an exten­sive review of Eisenhower’s intelligence needs and posits, on the basis of documents then

[2] The information about the Killian panel’s recommendations and the approach that Quarles made to members of the U. S. National Committee of the IGY comes from R. Cargill Hall’s article in the Quarterly of the National Archives.

[3] Robert Goddard launched the world’s first liquid-fuelled rocket on March 16, 1926. It reached an altitude of 41 feet and landed 184 feet from the launch site.

[4] Leonov told this story to Jim Harford, who is writing a biography of Sergei Korolev to be published by John Wiley and Sons in October 1997.

[5] Much of the information about Korolev’s early life is drawn from Yaroslav Golvanov’s book—Sergei Korolev: The Apprenticeship of a Space Pioneer.

[6] The words in quotation marks are extracted from Golovanov’s lengthier account, which appears in Sergei Korolev: The Apprenticeship of a Space Pioneer.

[7] In October 1953, President Eisenhower’s National Security Council endorsed a policy dubbed the “New Look.” This policy’s aim was that the United States should seek obvious strategic superiority and use rhetoric indicating a willingness to use it. The think­ing was that such a policy would deter Soviet aggression and return the diplomatic initiative (post Korea) to the U. S. and permit lower budgets. Eisenhower and his advisors had deter­mined that lower military spending was necessary because the levels at the time endangered national security as much as did inadequate arms. From. . . the Heavens and the Earth: a Polit­ical History of the Space Age, by Walter McDougall. Basic Books (1985).

[8] R. Cargill Hill points out that the president knew from intelligence gathered by the U2 spy planes that the Soviet Union did not have masses of ICBMs aimed at the U. S. The intelligence came from the illegal flights over Soviet territory, and Eisenhower told Secretary of State John Foster Dulles on the day before the telecast that he would not tell the nation that the United States had the ability to photograph the Soviet Union from high altitudes.

[9] At least two other people should be mentioned in connection with the early days of civilian weather satellites. They are Robert White and Fred Singer. White, who retired in 1995 as president of the National Academy of Engineering, was the head of the Weather Bureau and administrator of NOAA in the 1960s and 1970s. He was an influential sup­porter of satellite meteorology. Fred Singer, who was a member of the Upper Atmosphere Research Panel and worked for a while at the Applied Physics Laboratory, oversaw impor­tant engineering advances to TIROS. He was, says White, and is, a fascinating person and a maverick. His most recent provocation to the scientific community is a disbelief in the human contributions to global warming. In the early 1960s, several other people joined the new field of satellite meteorology, and anyone interested in learning about the technology in detail should see Margaret Eileen Courain’s Ph. D. thesis, Technology Reconciliation in the Remote Sensing Era of US Civilian Weather Forecasting, Rutgers University (1991).

[10] See RAND’s Role in the evolution of balloon and satellite observation systems and related US space technology, by Merton E. Davies and William R. Harris, published by the RAND Corporation.

[11] Technology Reconciliation in the Remote Sensing ERA of US Civilian Weather Forecasts, Courain, Rutgers University.

[12] Satellites in an orbit со-planar with the equator at an altitude of nearly 22,300 statute miles are called geostationary satellites because they can be regarded as stationary with respect to a particular point on the Earth, because the satellite takes about the same time to complete its orbit as the Earth takes to turn once on its axis. The satellite, however, is not exactly stationary with respect to the Earth because of the gravitational abnormalities caused by the Earth’s inhomogenous structure, but by the time of Suomi’s and Parent’s pro­posal, observations of satellites like Transit were beginning to improve the accuracy of grav­itational models, making it possible to predict shifts in the satellite’s position.

[13] The idea of taking advantage of a spinning satellite for an automatic east-west scan was not uniquely Suomi’s. A similar idea existed in the world of photo reconnaissance. Merton Davies, of the RAND Corporation, and Amron Katz had encouraged adaptation of a panoramic camera for high-altitude photography. Then Merton Davies had realized that the camera could be fixed to a spinning spacecraft to achieve an automatic east-west scan as a satellite, not in a geostationary orbit, traversed its orbit. From RAND’s Role in the Evolution of Balloon and Satellite Observation Systems and Related US Space Technology. Edwin Land showed the first satellite reconnaissance photographs from such a system to President Eisen­hower on August 25, 1960.

[14] In 1946, Louis Ridenour, an engineering professor at the University ol Pennsyl­vania, independently put forward the idea that a satellite in synchronous orbit would be a good place for a radio relay.

[15] The U. S. Air Force had an alternative approach to passive communication. This involved distributing five hundred million copper threads with a thickness one-third that of a human hair into orbits two thousand miles above the Earth. These would be spaced at five hundred-foot intervals and create an artificial ionosphere. It was not a popular idea with optical and radio astronomers. The scheme was initially called Project Needles, but because the name seemed too descriptive, it was changed to Project Westford. An attempt to distrib­ute the copper threads failed on October 21, 1961. In a research proposal prepared for inter­nal consumption, John Pierce and his colleague Rudi Kompfner said that Project Westford had very little to recommend it.

[16] Aerospace companies had their own ideas. Lockheed, for example, proposed that it, together with RCA and General Telephone and Electronics, should launch a system of spin-stabilized satellites into twenty-four-hour orbits. GTE had also had earlier discussions with the Hughes Aircraft Company.

[17] The story today has come almost full circle, and commercial plans for fleets ot medium altitude communication satellites pose a challenge to geostationary telecommuni­cation satellites.

[18] John Pierce knew them all: Arno A. Penzias and Robert W. Wilson, who discov­ered the cosmic background radiation using the horn antenna developed for the Echo spacecraft; and John Bardeen, William Shockley and Walter H. Brattain, who invented the transistor. Transistor, incidentally, is Pierce’s neologism. Walter Brattain asked him what to call the new device. Pierce writes in Signals: “I told him “transistors,” it seemed logical enough. There were already Bell system devices called thermistors, whose resistance changed with temperature, and varistors, whose resistance changed with current. I was used to the ring of those names. Also, at the time we thought of the early point-contact transistor (then nameless) as the dual of the vacuum tube; in the operation of the two devices the roles of current and voltage were interchanged. The reasoning was simple. Vacuum tubes have transconductance, resistance is the dual of conductance, and transresistance would be the dual of transconductance, hence the name transistor.”

[19] Bandwidth measures the frequency spread. John Rubels figures in a memo from DDR&E were that 20 words per minute in Morse code takes 9 cycles, 100 words per minute by teletype channel needs 75 cycles, voice telephone takes 3,500 cycles (3.5 kc), scrambled voice takes 50,000 cycles (50 kc), commercial TV takes 6,000 cycles (6 me). Today we would say Hertz rather than cycles.

[20] How clearly a signal is heard depends then on the power of the transmitter and the distance over which the signal is sent as well as on the signal-to-noise ratio and the bandwidth of the transmission. The noise, heard as static, comes from many sources. It might, for example, be interference from terrestrial transmissions, such as the Baltimore TV station that drowned out Transit M’s signal or the electromagnetic field generated by vibrat­ing electrons in the receiving circuitry.

The Bird’s-Eye View

It is obvious that in observing the weather through the “eye” of a high-altitude robot almost all the quantitative measurements usually associated with meteorology must fall by the wayside.

—From a Project RAND report: Inquiry into the feasibility of weather reconnaissance from a satellite (1951), by William Kellogg and Stanley Greenfield.

D

uring World War II, Japanese paper balloons floating on currents in the upper reaches of the lower atmosphere carried incendiary bombs across the Pacific to the United States. They caused some forest fires, which were quickly extinguished. Censorship kept news of the few fires from the public, and thus the balloons did not precipitate the widespread consternation that Japanese strategists had hoped for.

William Kellogg and Stanley Greenfield were intrigued by the story of these balloons and were impressed by the knowledge of the atmosphere that such a campaign had needed. The two men worked for a newly formed group of technical consultants known as the RAND Corpora­tion.[10] The interest the two men had in high-altitude balloons, which they believed might make good platforms for photo reconnaissance and intelli­gence gathering, evolved eventually into a conviction that satellites would provide a good platform from which to observe the weather. Their early conceptual work on “weather reconnaissance” became part of TIROS, and in 1960, the American Meteorological Society presented Kellogg and Greenfield with a special award.

The RAND Corporation was an ideal place for Kellogg’s and Greenfield’s work. The organization was formed in 1948 when the U. S. Air Force, previously the Army Air Forces, became a separate branch of the armed services. It grew from Project RAND, which the Douglas Cor­poration set up immediately after the Second World War to evaluate advanced technology for the Army Air Forces’ missions. RAND made its
first analysis of the technical feasibility of satellites for the Army Air Forces in 1946.

In 1948, the government assigned responsibility for satellites to the Air Force, which appointed RAND to manage the work. RAND subcon­tracted studies on guidance, stabilization, electronic reliability, communica­tions, space reconnaissance systems, and space power systems to companies such as Westinghouse, Bendix, and Allis Chalmers. In 1951, RAND pub­lished classified studies incorporating industry’s and its own work. These studies analyzed potential missions as well as engineering design, the polit­ical implications of the technology, and the potential of satellites for sci­ence. Later, when the U. S. space program got seriously underway follow­ing the launch of the first two Sputniks, the content of these reports would be incorporated into the early classified satellite programs, such as the Dis­coverer series (on which APL had a transmitter).

RAND’s main conclusion in 1951, however, was that satellites had potential as observation platforms for reconnaissance. In the same year, Kellogg and Greenfield completed a study on the feasibility of satellites— referred to as satellite missiles or satellite vehicles—for “weather reconnais­sance.”

Weather and reconnaissance satellites proved to be a hard sell, but of the two it was reconnaissance satellites that first won the administration’s backing. In winning that backing, reconnaissance satellites precipitated, by the tortuous paths that characterized the Cold War, the U. S. space pro­gram, including—eventually—meteorology satellites.

Initial opposition to reconnaissance satellites focused on their techni­cal limitations. In those days photography from high altitudes offered spa­tial resolutions of the order of hundreds of feet, knowledge obtained from high-altitude rocket photographs (thirty, forty-five, and sixty-five miles), which the Navy had been the first to take. This resolution was much poorer than what could be obtained from cameras on aircraft, and those who had developed expertise in photo interpretation during World War II were scornful of the technology’s capabilities.

But prompted by the need for better intelligence, both to prevent surprise attacks and to monitor arms-control agreements, the Eisenhower administration proposed low-level funding for a reconnaissance satellite program in fiscal year 1956 (that is, for funds available from October 1955). The money authorized was as follows: fiscal year 1956—$4.7 mil­lion; 1957—$13.9 million; 1958—$65.8 million.

Air Force historian R. Cargill Hall says that the critical factors in winning backing for the development of a reconnaissance satellite were that satellites offered a photographic platform that could not be seen by the naked eye or detected by radar sensors, and if satellites were detected, they would be too far away to be shot down.

These were not advantages that the Soviets were likely to appreciate. When President Eisenhower suggested in 1955 that there be an “open skies” policy, allowing overflight of Soviet and U. S. territory in order to ver­ify disarmament agreements, Khrushchev had dismissed the idea, calling it “licensed espionage.” Khrushchev viewed the policy as an attempt to gather information on potential military targets, and that, indeed, was one of its purposes. So the prospect of American spy satellites could be guaranteed to provoke Soviet animosity. For these and other reasons, argued the historian Walter McDougall, the Eisenhower administration wanted a civilian satel­lite launched first to establish the precedent of the freedom of space. In an article in Prologue, Quarterly of the National Archives (spring 1996), Cargill Hall confirms this view, arguing that the IGY enterprise effectively was made into a stalking horse for military reconnaissance satellites.

It was not a cheap stalking horse—nearly $20 million initially for the satellites alone—but it was effective. The program played host to enough political tensions, technical difficulties, protests from the Army, and criti­cisms of the Eisenhower administration’s Vanguard decision to keep the eyes of the world focussed on the IGY. Thus it was that IGY became the forum in which the first satellites were launched.

To the disappointment of Kellogg and Greenfield, weather satellites languished during the discussions about reconnaissance. Even though the resolution was poor, satellite images, they believed, offered something of potentially great importance to meteorologists—a “synoptic” picture, assembled from several individual photographs, showing cloud patterns over a large expanse of the earth at one time.

The proposal Kellogg and Greenfield made in 1951 had to wait seven years until the Advanced Research Projects Agency was prompted by Sputnik into finding applications for satellites. William Kellogg was then appointed to head an ARPA panel drawing up specifications for a meteorology satellite.

Kellogg turned to his and Greenfield’s early ideas, which, as in the case of reconnaissance satellites, owed much to rockets and high-altitude balloons. Images from rockets enabled meteorologists to begin developing analytical techniques that made sense of the bird’s-eye view of the Earth, while balloons served as test beds for meteorological instruments and cameras. The balloons could carry more weight aloft than could early satellites.

In the late 1940s, high-altitude balloons were already a versatile technology for a variety of applications. William Kellogg had worked on a project for the U. S. Atomic Energy Commission investigating their potential for monitoring the dispersion of radioactive particles from atomic tests; the Japanese, as we saw, had used them to carry incendiary bombs; and the Photo Reconnaissance Laboratory at the Wright Field in Ohio established that balloons made a stable platform for aerial photogra­phy.

At about the same time, in January 1949, the Bulletin of the American Meteorological Society published an article by Major D. L. Crowson, “Cloud Observations from Rockets.” Crowson suggested that even low-resolution imagery from high altitudes would improve weather forecasting.

Given this background, it is not surprising that Kellogg and Green­field decided to pursue the idea of using high-altitude balloons first for photo reconnaissance and then for meteorological research. This work gave them the information they needed about optical systems for their 1951 report on weather reconnaissance from a satellite. What they wrote was by no means a detailed engineering proposal, but it tackled concep­tually for the first time the elements of a meteorological satellite carrying a camera operating in the visible portion of the electromagnetic spec­trum.

They asked, Can enough be seen from an altitude of 350 miles to enable intelligent, usable weather observations to be made, and what can be determined from these observations?

From analyses of photographs taken by rockets, they decided that a resolution of five hundred feet was necessary if all the useful cloud struc­tures were to be identified.

They assumed that the camera would mechanically scan to build a photograph of a wide enough area—a 350-mile swath around the sub­satellite point—to be of use. Once they had specified the minimum reso­lution, they discussed the aperture, illumination, exposure time, and focal- length-to-aperture ratio needed to achieve a given contrast. The satellite, of course, would not be taking carefully posed and cunningly lit pho­tographs but would have to operate in whatever conditions nature de­creed. So the camera and optical system had to be chosen to provide usable photographs in a variety of conditions.

Photographs, they recognized, would not yield the quantitative information that meteorologists needed. It would be impossible to do more than make intelligent guesses at temperatures and pressures. But in those days, before numerical weather prediction, these limitations did not seem as great an obstacle as they would shortly become. Thus, unknow­ingly, Kellogg and Greenfield put their finger on what was to be the mam problem in winning acceptance for satellite meteorology. Analysts, they wrote, would have to make the most of the visible aspects of the weather in building their weather charts; clouds, being the most visible aspect of the weather, would be important.

Rocket photographs had already given an inkling of the inherent problems. In pictures taken by cameras on Thor and Atlas, it was difficult to tell whether areas of uniform greyness were cirrus clouds (wispy, high- altitude clouds formed of ice particles), tropospheric haze, or an artifact of the optical system resulting from the wavelengths accepted. All was not bad news, however, because more cloud patterns had been apparent in rocket photographs than had been expected.

In an attempt to get a feel for how accurately one might forecast weather from satellite photographs, Kellogg and Greenfield had estimated the synoptic situation from photographs taken during three rocket flights and had then made a forecast and compared it with records of actual weather on the day in question. Encouraged by the results, the two con­cluded that “combined with both theoretical knowledge and that gained through experience, an accurate cloud analysis can produce surprisingly good results.”

By the end of 1959, anticipating the launch of TIROS in 1960, lead­ing researchers met in Washington to discuss cloud research, a field known as nephology. Harry Wexler, then chief scientist of the Weather Bureau, pointed out that until the late nineteenth century, clouds had been almost the only source of information about conditions in the upper atmosphere, but that with the advent of balloon soundings, interest in nephology had declined. (Now clouds are recognized to be of crucial importance in meteorology, particularly in climate studies, and meteorologists are asking such basic questions as “What is a cloud?”)

Sig Fritz, who worked for Wexler, spoke of the strong sense researchers had that they would not know how to interpret cloud pho­tographs. This was a problem that Kellogg and Greenfield had foreseen in their 1951 report, and they had recommended that in preparation for satellite images, meteorologists build a comprehensive atlas of clouds as seen from above.

TIROS would soon begin that process.

The conception and birth of the TIROS satellites were difficult. First, in late 1957, the secretary of defense, Neil McElroy, agreed that a new agency—the Advanced Research Projects Agency (ARPA)—would have responsibility for key defense research and development projects. ARPA officially opened its doors on February 7, 1958, and weather satellites became one of its projects.

Immediately, Kellogg was asked to define the specifications for a weather satellite, which he did with the help of people like Dave Johnson. In the meantime, RCA had submitted a proposal to the Army Ballistic Missile Agency for a reconnaissance satellite as part of the Redstone mis­sile program.[11] The Department of Defense decided that the images from this satellite would not be good enough for intelligence gathering, and it therefore became TIROS, modified to incorporate the conclusions from Kellogg’s group at ARPA. ARPA managed the TIROS project until the National Aeronautics and Space Administration took it over in October 1958.

The newly formed NASA was a powerful organization, and it could expropriate groups and organizations. One of the groups that the agency wanted was that headed by Bill Stroud, of the Army’s Signal Corps of Engineers, which was working on a camera for the IGY’s cloud-cover experiment. After some altercations with military bureaucrats, Stroud was able to transfer to NASA in 1959, and he headed the agency’s meteorology branch at the Goddard Space Flight Center.

So TIROS had its roots in a spy satellite proposed to the Army by RCA but became a weather satellite managed first by ARPA and then by Stroud’s group at NASA.

The satellite’s optics had a field of view four hundred miles on a side. It carried a small Vidicon TV tube, selected because of its light weight.

The images recorded were poor because the camera’s electron-beam scan was not well controlled. Also, recalled Verner Suomi, who soon became involved with the TIROS program, “We didn’t know what the devil the damned thing was looking at. There were some problems as to what time the pictures were taken, and the spacecraft was spinning like a top. Where the devil was north? That caused major problems.”

Sean Twomi, of the Weather Bureau, soon identified the cause of one of the problems. The spacecraft needed to spin to remain stable in orbit, but the spin axis was tilting because of interactions between the spacecraft’s electrical systems and the earth’s magnetic field. Thus it was difficult to know where the cameras were pointing. Once Twomi had identified the problem, the Air Force developed magnetic stabilization to control the orientation of the spin axis, though it was some time before the engineer­ing problems of orientation and stabilization of weather satellites were fully solved.

Despite such difficulties, those involved, like the engineers and scien­tists at the Applied Physics Laboratory, had an overwhelming sense of being pioneers. Bob Ohckers, an engineer who worked on the TIROS program at RCA and later moved to Suomi’s lab at the University ofWis – consin, recalled, “In those days, there were no cutesy requirements, no quality control or oversight. Everything was experimental. If we had a fail­ure, we would try to keep it from the contractor, particularly from Thomas Haig, who headed the Air Force’s weather satellite program (Haig also joined the University ofWisconsin, where he and Suomi spent some time feuding). Haig would try to ferret out what was going on. We were told to tell him nothing. The whole group was working with rolled up sleeves and screwdrivers.”

Despite the limitations of the TIROS satellites, both in terms of the data they collected and of the analytical techniques available for data processing and interpretation, the first images returned to Earth were tantalizing.

In 1964 Suomi gave a lecture to children at a local school. A copy of the speech, recorded by some attentive listener, was in an old filing cabinet in the basement of the building where Suomi worked at the University ol Wisconsin. He told them how unsophisticated and crude the satellites were. But he also told them that during one orbit of TIROS they had identified two hurricanes in the southern hemisphere before they had been spotted by ships or weather stations. He showed them bright areas of cloud, telling them that this meant that the clouds were thick and high and represented an enormous thunderstorm, but that they only learned these things after the event. “We have much to learn about how to apply these pictures. The future depends not on the hardware, not on the gadgets, not on the software, but on individuals applying their knowledge to this very challenging problem [of interpretation].”

With the formation of NASA, the defense and civilian meteorology satellite programs went their separate ways. Until 1965, the Defense Mete­orology Satellite Program (DMSP) was one of a suite of satellites con­trolled ultimately by the CIA—an indication of how intertwined the mis­sions of reconnaissance and meteorology satellites were. Then, in 1965, control of funding for DMSP was transferred to the Air Force. DMSP remained secret until John McLucas, the under secretary for the Air Force, made the program public in 1973.

Since 1958, when the two programs diverged, the civilian and defense meteorological worlds have resisted all political efforts to reunite

them. Meetings about a merger between the two were, says one partici­pant, like arms control negotiations of old, where people developed their arguments as to why arms control was not possible.

Suomi, who like many of his colleagues worked on both sides of the fence, recalled that he would see the same work being done twice. “Part of my activities here [at the University of Wisconsin], which were classified

then, was to put the heat budget experiment on a military satellite. The thing that was interesting was that many of the things that the civilian pro­gram utilized were actually developed for the military. What interested me as a party to both was that I saw one part of the RCA building which was classified and I saw another part of the building not classified. Both parts were classed as development, but really they were only one development. Someone made a lot of money on that.”

Johnson, too, saw duplication, but said that interactions between the two worlds could work well when individuals in the military program were cooperative. On one occasion Johnson wanted to fly a new tube, and the DMSP allowed one of its own tubes to be replaced by Johnson’s. Many of the DMSP people would also, says Johnson, do what they could to move expertise from the “black” world. But technology was not always transferred. Suomi recalled that DMSP effectively equipped spacecraft

with “exposure meters” so that they could photograph clouds by moon­light in the visible part of the spectrum. Asked if such technology would have helped civilian weather satellites and why it was not transferred, he answered, “Yes” and “I don’t know.”

Though declassification may clarify some of these questions, it is unlikely to revise substantially the pioneering role that Verner Suomi played.

Chapter six: Heady Days

The contents of the National Security Council’s agenda and mention of Lay’s phone calls to Alan Waterman (page 58) are among Waterman’s papers in the Library of Congress.

Information about Tycho Brahe and Johannes Kepler (page 59) comes principally from an essay on Kepler by Sir Oliver Lodge in The World of Mathematics, edited by James R. Newton (Tempus, 1956). Textbooks con­sulted for pages 60 and 61 are Introduction to Space:The Science of Spaceflight, by Thomas D. Damon. A foundation series book, chapter three deals with orbits (Orbit Book Company, 1989); The Feynman Lectures on Physics, vol­ume one, chapter 7 (Addison Wesley, 1963).

Bill Guier and George Weiffenbach supplied the information for pages 61 to 65.

The textbook consulted for pages 65 to 69 is The Feynman Lectures on Physics, volume one, chapter 34 (Addison Wesley, 1963).

The material on pages 70 to 72 is based on the memories of George Weiffenbach, Bill Guier, and Henry Elliott. They all spoke to me inde­pendently. Guier and Weiffenbach spoke to me many times. Each remem­bered things a little differently, but their memories differed little in sub­stance. These memories are, to my knowledge, the only sources of information for Guier’s and Weiffenbach’s work in October at APL. As far as I know, there are no written records, not even laboratory notes, that support the assertion that Guier’s and Weiffenbach s work was unofficial and an indulgence of curiosity.

Keep it Simple, j| Suomi

We just got along like brothers. His interests and personality were what was missing from mine. I tend to dream a lot and jump to conclusions. He was thorough and patient. How hard we worked.

—Verner Suomi, talking about his friend and former colleague, Robert Parent

As I look back, I can see all the trial and error. I don’t want to use the word comedy.

—Leo Skille, an electronics technician who worked with Suomi and Parent from 1959

I

t rained on the morning of Verner Suomi’s memorial service. Deep – throated Midwestern thunder played its inimitable summer accompani­ment. And as one might expect among people gathered to celebrate Suomi’s life, there were questions—asked humorously, sadly—about just who it was that was responsible for the weather.

Ponytailed graduate students joined family, visiting scientists, and col­leagues from Suomi’s own academic home at the University ofWisconsin, Madison, in the Lutheran church on the edge of campus. Soon organ and trumpet released the opening notes of Giuseppe Torelli’s Sonata in D Major: lively notes that grew thoughtful, a little sad, and again lively.

Suomi’s son Stephen spoke of a man who could be stubborn, even too stubborn, but who would fake footprints and sled marks in the snow to convince him for one more year that Santa Claus existed. The minister spoke of a great scientist, but here he missed Suomi’s essential strengths, which lay in his engineering ingenuity, in his experimenter’s determina­tion to collect data, and in the stubbornness that his son recognized.

Suomi’s engineering ability, his “practical eye,” helped him to see how to build instruments that took advantage of the spatial and geometric relationships between orbits, the rotating Earth, and the satellites, which were themselves spinning. Time and again he watched as instruments

embodying his engineering concepts were launched. During interviews in the late spring of 1992, Suomi recalled that “what was exciting was to see this big rocket with ‘United States’ written on it, and your gadget was in the nose cone.”

It was Suomi’s stubbornness that placed the instruments in those rockets and kept his interest in satellite meteorology alive during the two decades it took him and other pioneers to find ways of extracting useful information from satellite data and to convince the wider meteorological community of the information’s validity. Verner Suomi held his own in the competitive and dynamic arena of meteorology. He knew quite well that he was stubborn, and said of himself, “I just yell and holler until I get my way.”

Building things was important to Suomi throughout his life. He was fascinated as a teenager by radio and flying. To his generation—the teenagers of the late 1920s—these comparatively new technologies were what computing is to the scientifically inclined youngsters of the late 1990s. It was almost obligatory to build your own radio set, which Suomi did during his summer holidays. He took a car mechanics course at school and worked on an aircraft engine as he trained to be a teacher at Winona State College in Minnesota.

As Suomi was contemplating these memories in his office, he recalled that some years past he had met an old girlfriend, someone from the days before he met his wife, Paula. This woman had commented on his patience with things but lack of patience with people.

“I thought a lot about that; it had a big affect on me and was a mile­stone. Since then, I’ve tried to be patient with people. Whether I was suc­cessful or not, well that’s another story The thing is, it’s easy to work with things, they don’t talk back.”

Given this aspect of Suomi’s character, it makes sense that at the age of twenty-seven he gave up teaching high school to take a degree in mete­orology at the University of Chicago.

Chicago’s meteorology department was run at that time by one of the giants in the history of the subject, Carl Gustav Rossby. Reportedly, it was after learning from Rossby of Lewis Richardson’s failed attempt to make a numerical weather prediction that John Von Neumann concluded that weather forecasting was an ideal problem for the new “high-speed” computers.

Suomi had learned a little about meteorology and weather forecast­ing while he was learning to fly but had dropped the subject after he had received his license. (As he talked about flying he wistfully remarked that he would like to fly “one of those big planes.”) At Chicago, he found Rossby to be an original and fascinating person, someone who “could talk you out of your shirt and make you think you had a bargain.”

Suomi also met Reid Bryson, who established a meteorology depart­ment at the University ofWisconsin, one that became among the best in the country. Bryson asked Suomi to join him in developing the depart­ment, which subsequently proved to be a source of discord between the two men. Suomi and Bryson developed a complicated relationship that incorporated friendship and professional rivalry. They had both moved to the University ofWisconsin in 1948, where Suomi completed the doctor­ate he had begun in Chicago, earning his degree at the age of thirty-eight in 1953.

In many ways, Bryson’s and Suomi’s rivalry typified the greater rivalry that existed in meteorology between atmospheric modelers and advocates of satellites. Suomi, of course, was the satellite advocate, inter­ested in improved weather forecasts; Bryson was more interested in com­puter modeling and climate. With these interests, both placed themselves out of what was then the mainstream of meteorology—Suomi because satellite instrumentation was not providing what meteorologists needed, and Bryson because in the late 1950s and early 1960s climate modeling and prediction were not quite respectable. In the pursuit of more accurate weather forecasts, both for long-range and short-term predictions, tension over resources and staffing arose between the two men. Neither was con­tent to rely only on traditional instruments and weather maps. Both satel­lite technology and bigger and better computers on which to run more detailed models cost money and call for expensive expertise, and local ten­sions mirrored these at a national level. The fierceness of the disagreement at all levels was fueled by genuine scientific disagreement about the correct way to move meteorology forward.

As a general rule, the modelers were theoreticians at heart, while the satellite advocates had the souls of experimentalists. Of course, experimen­talists had to understand theory and theorists had to incorporate the results of experiment, but there seemed and seems to be a difference between the two groups that can set the two sects at one another’s throats.

The difference was manifest between Bryson and Suomi. Bryson worked on theories of climate. Suomi wanted to design instruments to learn more about what were then the largely unexplored regions of the atmosphere in the Southern Hemisphere and at higher altitudes, regions which, in an interview with William Broad of the New York Times, Suomi called the ignorosphere.

In the end, of course, meteorology proved to be big enough for both of them and to need both of their approaches. And they continued to live as neighbors in the houses they had helped one another build.

During the 1950s, Suomi formed another important professional relationship, one that was synergistic. This was with Robert Parent, from the university’s Department of Electrical Engineering. Their first collabo­ration was in the design and construction of an experiment known irrev­erently as Suomi’s balls, and more properly by the title of their proposal to the International Geophysical Year, which was “satellite instrumentation for the measurement of the thermal radiation budget of the earth.”

This property of the earth—its thermal radiation budget—is a cru­cial aspect of meteorology. It may even be the most basic factor influencing the world’s climate. Thus, the radiation budget is of more interest to basic research than in the applied field of weather forecasting. And there is some irony in the fact that Suomi, who became known for designing satellite instruments for improving short-term weather forecasting and who had the differences of opinion that he did with Bryson, should have been the first to put forward a proposal to measure the earth’s radiation budget.

The radiation budget is the driving force behind both atmospheric and oceanic circulation. Very simply stated, it is the balance between the incoming solar radiation and the outgoing radiation from the earth. Over time and over the whole earth, the incoming and outgoing radiation streams balance one another. The importance of the radiation budget, however, lies not in this balance but in the way that the balance of the incoming and outgoing radiation varies with geographical location and time.

The outgoing radiation falls into two broad categories: that which is reflected from the atmosphere and the surface without any change in wavelength, and that which is absorbed and reradiated at longer wave­lengths from the surface and the atmosphere. The amount of radiation reflected by some structures, such as ocean, cloud, or icecap, is known as the albedo of that surface. Meteorologists know now that the mean albedo of the earth is thirty percent, but that individual surfaces can have albedos ranging from ninety percent for dry snow at high latitudes to fourteen percent for dark, moist soil. Since the albedo is greatest at the poles, and concomitantly, absorption and reradiation are lowest at these latitudes, heat flows to the poles from the equator, driving the atmospheric and oceanic circulation without which there would be no weather.

Many types of atmospheric behavior and processes connect the global radiation budget to the weather. They range from large-scale struc­tures, such as the circumpolar vortices that spawn the jet stream, down to familiar weather features like cold fronts and storms. So the radiation bud­get, while it is the reason for the existence of the weather, is not of imme­diate application to short-term forecasts. Nevertheless, it is crucial to any deep understanding of meteorology, to climate studies, and to the explo­ration of long-term weather fluctuations.

Satellites have provided values of the three radiation fluxes (incoming solar, upwelling reflected, and reradiated energy) that contribute to the earth’s radiation budget, and they have done so for different geographical locations and seasons. The next step in this field will be to disentangle the detail: How much of the reradiated energy comes from the surface as opposed to the atmosphere, for example, and how absorption and reradia­tion vary with altitude and the physical composition of the atmosphere.

In the late 1950s, much controversy still surrounded the basic ques­tion of the earth’s overall radiation budget. That controversy arose because scientists were calculating the radiation budget from too little data. They needed a Tycho Brahe, or several Brahes, who would collect the observa­tions “heaped on observation,” but before satellites, there was no practical way of collecting comprehensive data. Some sounding rockets and ground stations had measured values of, respectively, the incoming solar radiation and terrestrial radiation. And even before the administration announced the satellite program, the meteorology panel of the International Geophys­ical Year had plans to gather more ground-based observations to increase their understanding of the earth’s radiation budget. Then came the satellite program, with its offer of a new perspective on the earth, that made possi­ble Suomi’s proposal.

Suomi and Parent did not join the International Geophysical Year until nearly a year after President Eisenhower’s announcement that the United States would launch a satellite. Then, sometime in the spring of 1956, Suomi attended a lecture given by Joseph Kaplan, the “five cigar” chair of the U. S.’s National Committee—the man who had shepherded the U. S.’s IGY proposals past wary administration officials. Kaplan by now was promoting the IGY and the satellite program to a wider scientific audience and to the public.

As he listened to Kaplan’s lecture, Suomi knew that he wanted to participate in the satellite program. His doctoral thesis—about ten pages of text and thirty of charts—had described the radiation budget of a cornfield. That is, he was interested in what happens to the sun’s energy: what is absorbed, what reemitted, when and how. Such basic information is valuable to agriculture. And today such studies are also important for correlating satellite observations with what is actually happening on the ground.

Never niggardly with the scope of his ideas, Suomi thought he could study the earth’s radiation budget from a satellite, and that the project would be “much more interesting than the radiation budget of a lousy cornfield.”

After the lecture he spoke to Kaplan, who told him to contact Harry Wexler, then chief scientist of the Weather Bureau as well as the chair of the IGY’s meteorology committee and a member of the U. S. National Committee of IGY By now the satellite panel was assigning priority to the various payload proposals, and at this late stage the sponsorship that Wexler quickly provided was essential. “Harry rigged it up,” said Suomi, “so my little proposal snuck in about three days after the door was closed. I got about $75,000, which in those days was an enormous amount of money.”

Wexler and his colleague Sig Fritz worked on some of the theoretical aspects of the scientific consequences of Suomi’s instrument and told the satellite panel that he and Fritz were interested in exploring whether data from Suomi’s instruments would show a correlation between isolines of reflected radiation and large-scale weather. Wexler presented Suomi’s ideas on June 7, 1956. During the next six months he also solicited the support of other influential meteorologists.

Although Suomi and Parent did not submit a formal proposal until the beginning of 1957, their new idea upset the proverbial apple cart. It was not the last time that Suomi would disturb nicely laid plans with a last-minute idea. He was to do the same with his idea for a spin-scan camera.

In the case of the radiation budget experiment, the overturned apple cart belonged to Bill Stroud of the Signal Corps of Engineers. His idea, like Suomi’s, was too ambitious for the existing technology, but both were pushing the boundaries. And both saw conceptually the promise of the bird’s-eye view.

Stroud wanted to photograph clouds from orbit. He had two broad purposes: first, to seek ways to identify cloud types from above, and sec­ond, to estimate the albedo of clouds.

The panel was interested in Stroud’s idea but concerned about the difficulties of extracting any useful information from the photographs. This was the concern that Kellogg and Greenfield had expressed in their report of 1951 and that would surface again during the TIROS program. Nevertheless, by the time Wexler introduced Suomi’s idea, the panelists had decided to back Stroud’s work, though it was classified as one of the “priority B” experiments.

And that is the priority that the experiment retained until the turn of the year, when a flurry of events precipitated a face-off between Suomi and Stroud.

By the end of 1956, the panelists had money for only four experi­ments. One of these was to be the Naval Research Laboratory’s satellite designed to measure temperature and pressure in space, another would be Van Allen’s payload, the third was to be a geomagnetic experiment, and the fourth experiment was to be Suomi’s.

Stroud must have objected, because when the panel next met early in 1957, they had decided that he and Suomi should compete for the fourth launch vehicle and that at the end of 1957 the country’s leading meteorol­ogists should decide between them. The two raced to prepare their satel­lites. But then Sputnik I was launched, followed shortly by Sputnik II, and President Eisenhower’s administration brought in the Army. There were now enough launch vehicles, and all the meteorologists had to do was to endorse both project.

Stroud’s payload was launched first, on Vanguard II. His camera worked, but its images were of little use because the satellite was precessing wildly. It was not possible to say where the camera was pointing. Stroud’s problems sprang from the difficulties that arise with spinning bodies. The final stage of Vanguard II was spun so that it would be stable in flight. And the satellite was designed to spin up after separation from Vanguard so that it too would be stable in orbit.

When the IGY’s satellite panel met for the last time to wind up busi­ness, on July 21, 1959, John Townsend, of the Naval Research Laboratory,

Keep it Simple, j| Suomi

Verner Suomi posing with the Vanguard satellite that was launched on June 22, l959.The satellite ended up in the ocean.

gave a report of the Vanguard II launch that gives a flavor of the difficulties and the unknowns of rocketry and satellites in those days. One theory was that there had been a collision between the third stage and the satellite. The idea stemmed from what they had learned about Vanguard I, the launch that generated the data for O’Keefe’s deduction of the pear-shaped earth. In that case, the Vanguard team had asked Fred Whipple’s group of optical trackers at the Smithsonian Astrophysical Observatory to locate the third stage for them. Whipple’s group had found the rocket’s final stage to

Keep it Simple, j| Suomi

Robert Parent (left) andVerner Suomi work on the electronics of their radiation balance experiment.

be in a markedly different orbit from that of the satellite. “So,” reported Townsend, “even after what appears to be a shutdown [of the third stage], there is enough propulsion in that unit as it cooks, or burps, to account for several hundred feet per second difference between the bottle [i. e., the third stage] and the satellite itself.” In other words, there was enough dif­ference in velocity to send the two into slightly different orbits, raising the possibility of collision.

The alternative explanation centered around the fact that the spin up and separation might have occurred at the same time. This could have

Keep it Simple, j| Suomi

ExplorerVII which carried Verner Suomi’s and Robert Parent’s radiation balance experiment into space on October І З, 1959.The white hemisphere attached to the mirror is one of the bolometers that gathered data for the experiment.

caused a problem because of the third stage’s unburnt fuel. The bottle, Townsend told the group, was an unbalanced stovepipe with holes in it that was unpredictable mathematically. If the satellite’s spin thrusters fired while still attached, the bottle would precess wildly and impart that precession to the satellite. Whichever explanation was correct, Stroud’s experiment was in trouble because the satellite had no internal damping mechanism.

Six months later, on October 13, 1959, Suomi and Parent’s payload reached space. Suomi had a receiver in his bedroom. He received four sig­nals a day, two within 105 minutes of one another, then a similar set twelve hours later.

“We worked like dogs,” said Suomi of his and Parent’s work on the radiation balance experiment. Leo Skille, a lab technician who worked with Suomi and Parent, recalled that Suomi had an air mattress that leaked, and that after four hours Suomi would find himself awake on the hard floor. The mattress served as Suomi’s alarm clock, and it was not uncom­mon for Skille to come in at eight in the morning and see both Suomi and Parent still working.

Suomi spent most of his time in electrical engineering because the meteorology department did not consider the satellite work an enormous blessing, although it did not hold him back.

Suomi and Parent amused observers as, sporting the bow ties favored by some academics, they strode around campus deep in discussion. When they encountered a grassy quad, Parent would—so the story goes—walk around it, while Suomi strode diagonally across.

“I’m a person,” said Suomi, “who says ‘lets do this.’ It might be a crazy idea, but I don’t let initial problems stop me. Bob was there as an evaluator. He didn’t discourage me. He put a certain amount of reality into some of the wild things I thought about.”

“Suomi would go off at tangents,” said his technicians, “Parent would bring him back to Earth. Suomi would figure you could build an instru­ment as fast as he thought of it. We’d say,‘Watch out for Suomi on Monday mornings.’ He’s come in with a whole bunch of ideas he’d thought of over the weekend.”

The engineers would wonder sometimes what he was getting at, and they wondered whether it was just they that didn’t understand his idea or whether he was coming out of left field. He always wanted to know why something wouldn’t work. “He could be stubborn,” they said, “and there were big rows, but he never held a grudge.”

Suomi and Parent’s first satellite experiment comprised four ping- pong-sized balls on the end of Vanguard’s antennas—hence its irreverent nickname. One was white and thus reflected visible radiation but absorbed the longer-wavelength radiation reflected from the earth. One was black and absorbed all wavelengths, and the other two were shaded in such a way as to discriminate between incoming and upwelling terrestrial radiation. Each contained a resistor that enabled a determination of the balls’ tem­perature (the relationship between resistance and temperature for different types of material is well known). These numbers did not provide absolute values for the three different types of radiation, but they did show how the radiation balance varied.

Suomi and Parent did most of the early construction themselves. “We were probably as capable as any of the engineers,” said Suomi. The two of them built four models: one for mechanical and vibration tests, one for electrical tests, one flight model, and one backup.

The bane of their lives were the transistors, which they knew com­paratively little about because transistors were a new technology. Suomi would order them in batches of fifty, paying $35 for an npn type transistor and $65 for a pnp. (A $35 watch today, said Suomi, contains about five thousand transistors.) The transistors were supposed to work within cer­tain temperature ranges, but most of them clustered at the extremes of the stated range rather than at the center, as became usual when the technol­ogy matured. They tested each component by cooking it in Paula Suomi’s electric frying pan and then putting it in the freezer. Then they hooked each component up to an oscilloscope and checked its characteristics by observing the plots of voltage against time for various inputs. If it worked, it could be soldered onto a circuit board; if not, it went into the bin.

At each stage of the construction, there were laborious tests. “I can honestly say,” recalled Suomi, “that we never took any short cuts. If we made any changes, we made the tests. We almost worshipped that rule.

“I remember distinctly that as we worked on something, we would work it to death, and if the improvement was not satisfactory, we didn’t say we don’t have to worry about that anymore. We’d go and find a blackboard and try to understand the problem at a deeper level than just something we were soldering. For example, we needed a power amplifier to work over a wide range of temperatures, and so we needed transistors that would work over a wide range of temperatures. We tried to use silicon, but the only thing then that could handle enough power was germanium, but germa­nium is temperature sensitive; therefore we put in resistors to modify the power input. The resistors had to be reliable at a range of temperatures, so we put in thermistors to compensate for different temperatures.”

In the end, they had underestimated by a factor of a thousand the time needed for testing, but to Suomi’s irritation, the flight model was not put through the severest tests. It was felt that testing wore out the payload.

Their task grew more complicated when the Army became involved. All of those assigned launches were asked whether their payload could be modified for Explorer. Suomi and Parent were among those who built payloads for both types of satellite and launch vehicle. Werner von Braun told Suomi that if he could not have his experiment ready by the launch date, the rocket would carry a plaque reading “this space was reserved for the University ofWisconsin.”

Suomi and Parent’s first launch attempt was atop Vanguard on June 22,1959. That payload went into the ocean. When I asked whether he and Parent went for a drink after the launch failure, Suomi looked into the dis­tance for a moment and said, “Yes, we went for a drink.” He paused and added, “maybe more than one. We weren’t thinking yet about Explorer. It hurt too much to lose the experiment on Vanguard. I could tell you a lot of things we did and said then, but they’re unprintable.”

But they had to think about Explorer very soon, perhaps even before their hangovers cleared, because the launch was scheduled for July 16. Suomi and Parent observed this second launch from the blockhouse at Cape Canaveral. Photographers and journalists only a quarter of a mile away watched the liftoff and saw with horror that the rocket appeared to head straight toward them. The photographers continued shooting as those around them dived under cars and trucks.

The rocket was, in fact, veering toward a populated part of Florida, and the range safety officer deliberately exploded it. The blazing rocket fell within 150 feet of the blockhouse. Suomi and Parent were trapped for more than an hour. When they were allowed out, they went to the smol­dering rocket and hacksawed out their instrument package.

For the next launch, in October 1959, Suomi, who now saw himself as a jinx, stayed in Wisconsin. That rocket reached space, and the payload collected its data successfully.

The value of Suomi’s experiment is not that! it provided a global map of the radiation budget, because it did not. Rather it is in the existence of the idea and in the thought that went into exploiting this new place from which to make observations and in the painful lessons learned during the primitive days of the space age.

The first factor operating against Suomi was the low inclination of the orbit selected for Vanguard and later for Explorer. From the low inclina­tion the satellite could “see” only a limited area of the surface and missed the poles completely Secondly, the primitive state of computing and data processing imposed limitations. In June 1961, Suomi wrote to a science journalist that they had been drowned in data from Explorer VII. The satel­lite was supposed to have been turned off on October 13, 1960, but the timer failed, so it was still operating eighteen months later. Suomi wrote, “Records are now being corrected on a minimum scale—it’s reassuring to compare records a year apart… the condition of the black sensor is nearly the same as originally, but the white sensor has discontinued due to expo­sure to the sun’s radiation.”

Though the results fell far short of the ambitious plan, the first attempt had been made to gather comprehensive data about the most fun­damental process driving the earth’s atmosphere and oceans. And Verner Suomi was hooked by the new field of satellite meteorology. He went on to fly radiation budget experiments on two of the TIROS series of space­craft and on classified satellites. He later referred to results from all of these when he was arguing that the spin-scan camera, for which he is most well known, should fly on one of NASA’s experimental spacecraft—the appli­cation technology satellite, or ATS-1.

Chapter seven: Pursuit of Orbit

The main sources of information for this chapter are Bill Guier and George Weiffenbach.

The problems in putting this chapter together were that there are no primary written sources that directly confirm what Guier and Weiffen­bach did and when, and they have both told the story several times, including on tape in a 1992 APL video. Thus it took some time to recall memories.

The only way around this seemed to me to go over and over the same ground from as many different angles as possible. And both Guier and Weiffenbach seemed to take to this approach as the proverbial duck takes to water. Each time I gleaned another fact, no matter how small, from one of their colleagues or from a published paper, I went back to one or another of them to ask more detailed questions or the same questions in a different guise. As I learned a little more about the physics for myself, I also went back to them.

The result is chapter seven, which is corroborated as much as possible by memories from other people.

Certain bounds to the time when their work was done are set by undis­puted dates, such as the launch of Sputnik II and the day that their work became an official project.

Lee Pryor, who was at that time studying computing at Pennsylvania State University, confirms much of what Guier says about coding for the Univac.

The richness of information available on the Doppler curve (page 75) is apparent in a highly mathematical in-house paper (Part of APL’s Bumble­bee series) “Theoretical Analysis of the Doppler Radio Signals from Earth Satellites” published in April 1958.

Charles Bitterli remembers working on an algorithm for least squares (page 78).

Henry Elliott’s memories corroborate Weiffenbach’s view of himself as a painstaking researcher who would check the quality of data in detail (page 79).

Chapter eight: From Sputnik II to Transit

Project D-54, to determine a satellite orbit from Doppler data, APL archives (page 82).

Guier and Weiffenbach’s briefing about their work (page 82) is from my interview with Harold Black.

Information about Guier’s and Weiffenbach’s early work on the ionos­phere (page 82) is from interviews with Weiffenbach and Guier.

A textbook consulted on ionospheric refraction is The Feynman Lectures on Physics, volume one, chapter 28 (Addison Wesley, 1963).

Weiffenbach’s memo to Richard Kershner and the first Transit proposal (pages 84 and 85) are in the archives of the Applied Physics Laboratory

Henry Riblet told me of the need to modify the design of circularly polarized transmitters for Transit’s spherical surface (page 85).

Information about O’Keefe and his views (pages 85 and 86) came from my interview with John O’Keefe.

Views about Frank McClure’s character (page 87) came from nearly every member of the Transit team that I interviewed.

Frank McClure’s ideas for a navigation satellite are in a memo dated March 18, 1958, reproduced in The First 40Years, JFIUAPL (Johns Hop­kins University Press).

What McClure said to Guier and Weiffenbach about his satellite naviga­tion idea is a story that both told me separately (page 88).

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.