Category Something New Under the Sun

The Realities of Space Exploration

We were pioneers, and we knew it.

—Bill Guier

P

arsons auditorium was crowded. Everyone was eager to hear the news as it was relayed from the Cape. They knew about the delays that had accumulated during the final countdown, heard the announcement to switch off radio frequency generators at the lab. The moments before a launch are always tense. In the final seconds the tension was alleviated, as the voice from the Cape intoned, “twelve, eleven, ten, eight, whoops, seven, six, five, four, three, two, one.” The Thor-Able rocket lifted off, car­rying Transit 1A aloft. They knew that Air Force radars were tracking its ascent; that engineers were calculating position, cross checking their slide – rule calculations and sending course corrections to the launch vehicle as needed. They heard the satellite’s transmitters and knew that everything was going well.

Then the transmitters stopped. For a while, no one knew what was happening. Then came the news that the third stage had failed. In all prob­ability, it and the satellite burned up on reentry into the atmosphere some­where over the North Atlantic, west of Ireland. Lee DuBois, one of the mechanical engineers, looked around the room. He saw the tears of disap­pointment on his colleagues’ faces.

The progress reports that APL sent to ARPA were as emotionless as those that described the shattering of their best satellite the previous month. The launch failure, it seemed, could be ascribed to the retro rock­ets on the second stage. These rockets were supposed to slow the second stage after separation of the third, so that the second stage would not inter­fere with the third as it coasted away prior to firing its own engine. When the retrorockets failed, the second stage bumped into the third stage, dis­rupting the third stage’s ignition sequence.

Transit 1A’s flight had lasted twenty-five minutes. Its electronics had survived the launch. As soon as the nose fairing that protected the satellite on liftoff had peeled away, all four frequencies were transmitted. The lab
immediately started an analysis of the telemetry, which comprised mea­surements of variables such as the satellite’s temperatures and solar cell voltages.

APL also had some Doppler data from the short period before the signal was lost. They were incomplete; Henry Elliott’s record shows that one signal was lost intermittently. For a while, an operator had locked unwittingly onto some other unknown signal. Signals from a TV station in Baltimore interfered with reception a few minutes later, and halfway through the pass, one of the tracking filters lost its lock. Nevertheless, APL learned enough to confirm “at least partially” that the ground stations’ design and operation worked, according to the progress report.

With the data received the computing team also made a rough cor­rection for ionospheric refraction. Then they set themselves a theoretical problem, imagining that the Doppler data from Transit’s brief sojourn in space had, in fact, come from a satellite in orbit. They attempted their least squares fit. Though they clearly could not check the accuracy of their “orbital determination” against prediction of the satellite’s position during its next orbit, they could check their results against the Air Force’s radar data. They found the least squares fit was closer than it had been for the Sputniks and Explorer 1 and were encouraged. Thus, though the failed launch did not yield what they had hoped for and pointed to problems that needed to be addressed, the team did learn some things.

For a definitive analysis of ionospheric corrections and to begin investigating the earth’s gravitational field they needed a successful launch. The next attempt was set for April 13, 1960. By January of that year, Kershner was coordinating preparations for Transit IB’s launch and that ol Transit 2 A. Transit IB would be similar to the lost satellite, but Transit 2A, scheduled for a June launch, would test different aspects of the proposed navigation system.

Details piled on details. All over the United States, presumably in the USSR as well, teams of engineers and scientists were slowly coming to terms with the complexities of space exploration. Memos in English and Russian were written, which, if they were like Kershner’s, covered an array of newly recognized problems that now are familiar to those in the space business: nose fairing insulation, loads on structures, details about an epoxy bond, maximum satellite skin temperature at launch, radio frequency links, concerns about deflection and vibration characteristics of the launch vehi­cle’s second stage, and on and on.

Simultaneously, preparations were going forward for the satellites that would follow IB and 2 A in the Transit experimental series with the physics, the engineering, and computing all being developed in parallel— at a time when computing and space exploration were new.

At the ground stations, repeated preparations were made during the first three months of 1960 to track one of the Advanced Research Project Agency’s Discoverer satellites, which was carrying a Transit oscillator (ToD Soc Transit on Discoverer).Transit on Discoverer was part of a program to develop precision tracking for reconnaissance satellites, and the launch was postponed repeatedly. The postponements complicated preparations for IB, as did expansion of the Transit control center and its communication links to encompass the other agencies that were now interested in the project and its data, including NASA and the Smithsonian Astrophysical Observatory.

During the same period, Kershner lined up the Naval Ordnance Laboratory to do magnetic measurements and experiments. The Transit team was interested in fitting its satellites with magnets to stop them from spinning (de-spin, in the industry’s jargon), control their attitude, and pro­vide stabilization.

By March, B and 2A were in the final stages of fabrication or test­ing. John Hamblen (who was Harry Zinc’s and Henry Elliott’s boss) decided that some discipline was needed. He had found out that flight hardware had been released before necessary electrical and environmental tests had been run. In a casually typed note he asked that in future those fabricating the satellite proceed to incorporate a component only if an engineer had first signed the test data sheet. Verbal assurances about a par­ticular component, he wrote, would not do. Thus, casually, at APL and doubtless in many other labs, was the need for documentation recognized, documentation that now, assert many engineers and managers, has grown out of proportion to its usefulness.

The year advanced to Wednesday, April 13, 1960. That was a long day at APL. The launch was scheduled for 7:02 A. M. Eastern Standard Time. Once again Parson’s auditorium grew crowded. Probably the room looked as it does in photographs of the launch of Transit 4B. The ashtray filled to overflowing on a table crowded with papers. Gibson, Kershner, and Newton formal in dark suits, others in shirt sleeves. Gib­son standing, pipe in hand. Kershner in headphones, or telephone to one ear, hand covering the other. Newton seated, twisted slightly to view over his shoulder the clock held at eight minutes to launch, frowning, as was Kershner.

For Transit IB, the countdown proceeded. The voice over the inter­com from the Cape would have been saying things like, programmer starts … gyros uncaged… electrical umbilical ejects… lift off (at 7:03 A. M.). But it was not yet time for the champagne. The satellite still had to reach orbit, which it did, though barely Instead of the nominal 500 nautical mile circular orbit, IB went into an orbit with a perigee of 373 nautical miles and an apogee of 748 nautical miles. Such a result was very inaccurate by today’s standards, but more precise orbits had to wait until those designing launch vehicles were able to perfect inertial guidance controls.

Transit IB’s orbit was, however, sufficient to allow APL to begin work checking whether two frequencies would be adequate to correct for ionospheric refraction or whether a greater number would significantly improve the correction. The answer was that two seemed to be sufficient, though more remained to be done before this question was finally settled.

The immediate task on the first day was to determine an orbit, then to predict its position for the next twelve hours. Until midafternoon, there were computer problems. Then at 15:30 they determined their first orbit. The curves did not fit well, but they thought that this might be because the satellite was still spinning. Spinning ceased on April 19. On April 20, they determined another orbit from observations of fifteen passes at five locations. Again there was a poor fit. They decided this time that the prob­lem was noise. Like Transit 1A, Transit IB carried four frequencies for the investigation of ionospheric effects. Now they turned their attention to the second frequency pair, and the fit was better.

With the data from the second frequency pair, they determined satel­lite position to within 150 to 200 feet from observations of a single pass over a limited region of the earth. With data for half a day from the differ­ent tracking stations they could, assuming a simplistic model for the gravi­tational field and uniform air drag, determine satellite position to within one nautical mile. The longer the arc, the poorer the accuracy appeared to be. Something seemed wrong. Over and over again they looked for errors in the data and software. They could find none. It was a troubling situa­tion.

Extrapolating from a day’s observations, they then predicted the fol­lowing day’s orbit. This was what it was all about, developing a way of pre­dicting an orbit so that its coordinates could be uploaded to the Transit satellites twice a day, enabling the submarines to fix position with respect to a satellite in a known position.

The Transit team looked for the satellite at the time and location they had predicted.

And then they knew they were in trouble.

There was a discrepancy of two to three miles between prediction and observation. While this much error had been acceptable when they were first establishing Sputnik’s orbit in the fall of 1957, it was unaccept­able as the basis for a navigation system. “The satellite,” recalls Guier,“was all over the sky.” Again, they thought that it was a problem with the pro­gramming. But it wasn’t. What they had suspected but had not fully rec­ognized, and what O’Keefe had repeatedly warned Guier and Weiffen – bach about, now came to dominate the theoretical analysis of satellite motion. Earth’s gravitational field was far more complicated than anyone then knew. O’Keefe, because he knew about the perturbations in the moon’s orbit, was expecting that satellites in near-Earth orbits would show more pronounced perturbations, but even he could not have antici­pated the huge variation and the complexity of the gravitational field that was to emerge.

For the position fixing accuracies they wanted to achieve, their knowledge of the gravitational forces perturbing near-Earth orbits needed to improve considerably.

There were precedents. Others had wrestled with apparently unruly satellites. Not least of these were the men within whose paradigms early satellite geodesists were working—-Johannes Kepler and Isaac Newton. Both had struggled to understand the nature of orbits as, mystics both, they sought glimpses of fundamental truths about the universe. Kepler’s focus was on the sun’s satellite Mars; Newton’s was on the earth’s moon. In his book The Great Mathematicians, Henry Westren Turnbull writes, “The Moon, for instance, that refuses to go round the Earth in an exact ellipse, but has all sorts of fanciful little excursions of her own—the Moon was very trying to Isaac Newton.”

And very trying would be the motion of satellites in near-Earth orbits to the early satellite geodesists who, with the technology to observe satellite motion in greater detail than could Kepler or Newton, noticed a veritable plethora of fanciful excursions. The forces causing these devi­ations from elliptical motion needed to be accounted for so that their effect on satellite motion could be quantified and thus orbital prediction improved. It turned out also that because the irregularities in the gravita­tional field are due to variations in the Earth’s shape and composition, sci­entists reaped an unexpected and abundant scientific harvest from observa­tions of orbits. Satellite geodesy supplied, for example, some of the evidence for the theory of continental drift and thus for theories like plate tectonics.

APL was one of the early groups observing satellite motion. They were impelled by the unlikelihood that other geodesy programs would meet Transit’s needs by the time the system was scheduled to be opera­tional, at the end of 1962.

Like other satellite geodesists around the world, the Transit team wanted to determine the “figure” of the Earth. The Earth’s figure is not the topography that we see; rather it is a surface of equal gravitational potential (a geoid) that coincides with mean sea level as it would be if the sea could stretch under the continents. This geoid looks like a contour map. It has highs and lows that represent how the gravitation potential differs at a particular geographical location from what the potential would be at that point if the earth were a water-covered, radially symmetrical rotating spheroid (an ellipsoid of revolution), not subject to the gravita­tional pulls that cause tides. This hypothetical surface is known as the ref­erence ellipsoid.

It is the differences in gravitational potential between the figure of the Earth and the reference ellipsoid that geodesists study as they seek clues to the earth’s shape and structure. At first, only the deviations in motion caused by large irregularities, such as the pear-shaped Earth, were included in geoid models. Today’s models include the gravitational conse­quences of localized irregularities in shape or density. In the mid 1990s, the most accurate geoid maps available to civilians were of what is termed “degree and order 70.” Generally speaking, the higher a model is in degree and order, the more detailed is its description of gravitational potentials, in much the same way as a finer scaled topographical map gives greater detail about a piece of terrain. A geoid map, however, cannot be understood by analogy to an ordinary map. The gravitational potential at a given location is attributable not only to the local features, but also to the varying lengths of gravitational pull exerted by everything else. And the higher the degree and order of a geoid map, the more geologists can infer about the Earth’s structure. Geodesists aspire in the next century to satellite-based models that will be accurate to degree and order greater than 300, the goal being

to provide data that will help geophysicists to understand the earth’s geo­logical origins and history

The road to such comprehensive understanding of our Earth opened with the launch of Sputnik 1. Prior to the advent of satellites, geoid maps showed modest highs and lows that were a result of local measurements of gravity. The force of gravity exerted on a satellite’s motion, though, includes the sum of all the gravitational anomalies resulting from every irregularity of shape and density in the Earth. Disentangling these effects and relating them back to a specific aspect of the earth’s physical nature is a little like unscrambling an egg. Nevertheless, with extensive computer modeling the job can be done.

APL produced the first American satellite geodesy map in 1960, a crude affair by comparison with those of today. Guier and Newton led this effort and found that as with orbital determination and satellite navigation, they had again provoked hostility. Their early geoid maps showed far greater highs and low than appeared in maps from presatellite days, and traditional geodesists dismissed them as amateurs.

APL continued to produce geoid maps of increasing sophistication, but much of this work was classified. Civilian scientists at places like the Smithsonian Astrophysical Observatory and the Goddard Space Flight Center soon came to dominate the field, though APL’s work filtered dis­creetly and obliquely along some grapevines.

The lab’s first gravitational model contained a value for the Earth’s oblateness that was more accurate than that existing pre satellites as well as a term describing the pear-shaped Earth. Shortly afterwards Robert New­ton at APL and independently the Smithsonian Astrophysical Observatory made the next big discovery, which was that the Earth is not rotationally symmetric about its axis. Just as the northern and southern hemispheres are asymmetrical, so too were the eastern and western hemispheres. A number of scientists, most particularly the Soviets, had suspected that this might be true. Later on, APL optimized their geoid maps for Transit’s orbit; that is, they only unscrambled those aspects of the egg that affected polar orbits at Transit’s altitude.

The principle involved in extracting information about the Earth from satellite data is simple to explain in general terms, but very difficult to apply in practice: observe the satellite, note its departure from elliptical motion—its “fanciful excursions”—and try to find (in the computer model) what aspect of the Earth’s shape and structure, for example a par­

ticular dense structure or a liquid area, would give rise to the gravitational anomalies that would cause the satellite’s observed departure from an ellipse.

More detailed gravitational models and ionospheric corrections enabled the orbital determination group to improve their knowledge of satellite position from between two and three kilometers in 1959 to a little under one hundred meters by the end of 1964. With problems like iono­spheric refraction corrected for, other problems emerged. Would it be nec­essary, they wondered, to correct for refraction in the lower atmosphere? Such refraction was a source of bias in their data that could make the satel­lite appear to be about half a nautical mile away from its actual position. Helen Hopfield, whose dignified presence could reduce unruly Transit meetings to silence, tackled this problem, and APL made corrections for tropospheric refraction.

When the Transit group compensated for motion of the geographic poles from their mean position in the early 1970s, the satellite’s position was known to within twenty-seven feet. Polar motion, caused by preces­sion of the earth’s spin axis due to the earth’s nonuniform shape and struc­ture, changes the position of a ground station by about a hundred feet per year, thus introducing a small error into the orbital determination and prediction. The error was negligible for navigators but important to sur­veyors.

By the time of the first launch, APL had stopped characterizing orbits solely in terms of Kepler’s elements (as had other groups). First, because motion within a single orbit does not exactly obey Kepler’s sec­ond law—there are small deviations, and the elements are actually average values. Second, even these average values change gradually as the orbit shifts in inertial space because of the gravitational consequences of physical irregularities in the Earth. For navigation and geodesy, averages were not good enough. It was necessary to know as exact a position as possible at given times in the orbit.

So satellite position was expressed in terms of Cartesian coordinates centered on the earth’s center of mass, with one axis aligned with the earth’s spin axis and the other two lying in the Earth’s equatorial plane. The orbital prediction was made by finding the acceleration from New­ton’s second law of motion—the famous F = та that is so crucial to sci­ence and engineering, where F is force, m is mass, and a is acceleration. The value of the force acting at different parts of the orbit comes from the model of gravitation; then numerical integration of the components of acceleration, a = F/m, yields position and velocity at any desired instant of time.

If it had not been for the new generation of computers, typified by the IBM 7090, this work would not have been possible. The 7090 was one of the newest and best when it was installed in August 1960. It could per­form 42,000 additions and subtractions per second and 5000 multiplica­tions and divisions per second, and it could store 32,768 words (approxi­mately 0.03 megabytes). The 7090 was almost fully transistorized, unlike the vacuum-tube Univac.

The Univac had been badly stressed by the orbital determination program, taking eight hours for eight hours worth of prediction. The IBM 7090 could do the same job in an hour. To run the early gravitation mod­els on the Univac, which embodied only a few of the terms representing the earth’s gravitational field, Guier would set aside three or four week­ends. Had the Univac, which contained vacuum tubes with a mean time to failure of between 15 and 20 hours, been called upon to run the gravita­tional models that were to appear in the coming few years, it would undoubtedly have broken down. Even the 7090 would soon have to be updated as the gravitational model grew more intricate. Today Cray super­computers run some of the largest models; desktop machines with Pen­tium or 486 chips can run models of degree and order 50.

During 1960, Newton, Guier, Black, Hook, and others prepared for the transition to the 7090. The programs had to be rewritten in an assem­bly language compatible with the 7090’s architecture. The orbital determi­nation program occupied four or five trays of punch cards. Woe betide the person who dropped one. And drop them they did, recalls Black, with a laugh that has an edge even after thirty years.

Black and his colleagues were also learning—painfully—about soft­ware engineering, a nascent, scarcely recognized field. Black’s job was to get the orbital determination program running. He was starting with the physics developed by scientists like Newton and Guier. They generated the equations representing the physical realities, and as they understood more about what was going on they generated more equations. Black learned early to freeze the program design and fold new equations repre­senting the physicists’ deepening understanding of the situation into the orbital determination program in an orderly fashion rather than piece­meal. That, at least, was his aim; but Black’s position between the scientists

and the programmers who wrote and tested the code was at times unenvi­able. He had to force agreement out of the scientists, and he fought Guier (his immediate boss), Newton, and Kershner, telling them, “You ain’t gonna change this damn thing.”

In 1962, Lee Pryor, who retired in 1995 as the last project manager of Transit, arrived at the lab. Pryor had specialized in computing while tak­ing his degree in mathematics at Pennsylvania State University. His first three months at college were spent writing programs in anticipation of the arrival of Penn State’s first computer. Black says that Pryor was a godsend. When he arrived at APL, the lab was putting the finishing touches to the first “operational candidate” of the orbital determination program. “We just needed to get it out the door,” recalls Pryor.

In 1962, much physics and mathematics remained to occupy the Transit scientists, but the computing was moving from their purview to that of the professional programmers like Pryor who were writing code for an operational situation rather than for research. The move was neces­sary because, while the scientists could write programs for their own research needs, their programs, it seems, could be cumbersome and prone to breakdown in operation.

Once work on the gravitational model was well in hand, it became apparent that the effects of air drag and the pressure of radiation from the sun would have to be considered. These were dealt with in the 1970s pri­marily by an elegant piece of engineering invented by Daniel De Bra from Stanford University. The navigation satellites were placed inside a second satellite. The separation between their faces was tiny. Sensors on the Tran­sit satellite detected when the inner satellite moved toward the outer sur­face, and tiny rockets moved the inner satellite to compensate for these forces, before they could offset the Doppler shift. An engineering solution was necessary because the time, size, and place of the forces could not be predicted.

In the mid 1960s, the failure of solar cells threatened the reliability of the operational Transit satellites. Until this problem was solved with input from Robert Fischell, the Transit satellites tended to fail within a year of launch. Once solved, some veteran satellites exceeded twenty years in operation. The Transit team also launched the first satellite with gravity – gradient stabilization, in which an extended boom encourages the satellite to align itself with the earth’s gravitational field. APL’s first—unsuccess­ful—attempt with this technology was on a satellite known as TRAAC,

The Realities of Space Exploration

Doppler shift due to satellite pass.

which also carried instruments to explore and characterize the Van Allen radiation belts. Ironically, the satellite failed because of ionized particles created artificially by a high-altitude nuclear explosion—as did many other satellites.

TRAAC carried a poem engraved on one of the satellite’s instru­ments. It was written by Thomas Bergen, of Yale University and is reprinted at the end of this chapter. Its mixture of hubris and wistfulness captures something of the atmosphere that surrounded the early work on satellites.

In the case of APL, that work led, of course, to the Navy’s Transit Navigation Satellite System. The lab built the experimental series, the pro­totypes, and many of the early operational satellites. For a time, Navy Avionics built some operational satellites, but the job reverted to APL when these proved unreliable. Eventually, RCA won the commercial con­tract for construction. More satellites were ordered than were needed,

because a problem with the solar cells that was reducing their operational life was solved after the contract was placed.

During the 1980s, under Bob Danchik’s tenure as project manager, when GPS was nipping at Transit’s heels, these satellites were finally launched. The last Transit satellite went into orbit in 1988.

Although there was always at least one operational Transit aloft and available for the submarines from 1964 onwards, the system was not declared fully operational until 1968. At that time four satellites provided global coverage. Not the instantaneous, precise three-dimensional position fix offered by the twenty-four-satellite constellation of GPS, but still, for the first time, an all-weather, global navigation system, a system developed initially for the military, but which evolved until ninety percent of its users were civilian.

In ten years, a newly perceived consequence of the Doppler effect in the three-dimensional world grew through all the stages necessary to design and engineer a space-based navigation system. The program began at a time when vacuum technology was giving way to transistors, when programs were written laboriously in assembly language, and when no one knew how to develop large software applications. The conditions in space were unknown. The physics of the newly entered environment had to be analyzed theoretically and understood experimentally. The complex nature of the earth’s gravitational field had to be researched and a provoca­tive new understanding of the geoid developed. Launch vehicles were imprecise in their placement of satellites, if the satellites reached orbit at all. Satellite design was a new field, with stabilization, attitude control, and communication between space and Earth all unknowns.

During the early development of Transit, the launch vehicle changed, the computers changed, and programs had to be rewritten. Ground stations and satellite test facilities were built. Programs and equip­ment had to be developed for the submarines. It is hardly surprising that one or two people say that they ended up in the hospital, nor that the effort is remembered vividly and affectionately. But as Pryor noted shortly before he retired in 1995, it was time for the Transit program to end.

When the Navy switched off the last Transit satellite in early 1997, it ended the longest-running singly-focused space program to date. It sev­ered the last direct link with the opening of the space age, closing the doors on that shed on the plains of Kazakhstan and on the cold morning when Sergei Korolev thanked his exhausted and elated engineers who had launched Sputnik I, which Guier and Weiffenbach would track, providing the basis for Transit, which helped Polaris, America’s riposte against the Soviet threat of nuclear attack, firing the rockets with whose develop­ment Korolev had been so involved, because he believed that rockets were defense and science, which they became, for both sides, as did Tran­sit, which also became important to civilians. Here is one thread in the Cold War.

And one wonders.

What would have happened if McClure, say, or Pickering, Milton Rosen, or von Braun had met Sergei Korolev? If they had been in a room with chalk, blackboard, and a problem? Faintly, one hears the voices, dis­cerns in imagination the energy and the imminent verbal explosions as Korolev’s little finger lifts toward his eyebrow….

For a Space Prober

by Thomas G. Bergen

From time’s obscure beginning, the Olympians Have, moved by pity; anger; sometimes mirth, Poured an abundant store of missiles down On the resigned defenseless sons of Earth.

Hailstones and chiding thunderclaps of Jove, Remote directives from the constellations:

Aye, the celestials have swooped down themselves, Grim bent on miracles or incarnations.

Earth and her offspring patiently endured, (Having no choice) and as the years rolled by In trial and toil prepared their counterstroke— And now tis man who dares assault the sky.

Fear not immortals, we forgive your faults,

And as we come to claim our promised place Aim only to repay the good you gave And warm with human love the chill of space.

 

Chapter three: Follow That Moon

William Pickering’s state of mind and actions following Lloyd Berkner’s toast to the Soviets come from my interview with him. He described also the error in calculation they had made and the phone calls that poured into the headquarters of the IGY (pages 30 — 34).

Information about Project Moonwatch comes from my interviews with Roger Harvey, Henry Fliegel, and Florence Hazeltine.

Information on the radio tracking program comes from interviews by Green and Lomask with Daniel Mazur and Joseph Siry in the NASA His­tory Office, as well as from the following papers: John T. Mengel, “Track­ing the Earth Satellite, and Data Transmission by Radio,” Proceedings of the IRE (44), 6,June 1956;John T. Mengel and Paul Hergert, “Tracking Satellites by Radio,” Scientific American (198), 1, January 1958.

Information about the goals of the IGY satellite program and details of the optical and radio tracking systems and the technical and budgetary difficulties faced comes from minutes of the IGY committees, subcom­mittees, panels, and working groups:

Minutes of the first meeting of the Technical Panel on the Earth Satellite program (TPESP), October 20, 1955. At this meeting the panel defined the program’s goals (page 32).

10 November 1955: An ad hoc meeting of the technical panel on Earth satellites (TPESP) convened to discuss the budget for the program, which had to be ready for a presentation to Congress and the Bureau of the Budget (predecessor to the current Office of Management and Budget) by March 1956. Homer Newell said that important things to be budgeted for were radio and optical tracking and scientific instrumentation. The NRL, who were the experts at radio tracking, wanted stations distributed between latitudes of 35 degrees north and south of the equator. The TPESP wanted to add two more tracking stations to extend coverage to 45 degrees. These tracking stations eventually became known as mini­track.

The optical tracking program was discussed in greater detail at the second meeting of the TPESP, on November 21, 1955. Fred Whipple, director of the Smithsonian Astrophysical Observatory, presented a report prepared by himself and Layman Spitzer. The TPESP recommended that up to $50,000 be awarded to the SAO immediately to set up a series of observ­ing stations. At the time, Whipple’s proposal was for twelve observing sta­tions and an administrative and computer analysis center. He also called for collaboration with amateur observers.

During the third meeting of the TPESP, on January 28, 1956, the difficul­ties of tracking began to emerge. A letter from Homer Newell on the problems of visual and photographic tracking of Earth satellites was read. It was not known whether radio tracking would work (see page 36). The expectation at the time was that there was only a fifty percent likelihood of minitrack succeeding; hence the need for optical tracking.

27 June 1957: The twelfth meeting of the USNC pointed out that there were still problems with the tracking system.

At the seventh meeting of the TPESP on September 5, 1956, John Hagan and Fred Whipple respectively updated the panel on radio and optical tracking. By now, Whipple had made contact with amateurs in an attempt to improve the chances of acquiring the satellite optically. The army, for example, had four hundred binocular elbow telescopes that volunteers, like Florence Hazeltine, could use at military bases.

The twelfth meeting of the TPESP, on October 3, 1957, the eve of the launch of Sputnik, opened with a discussion about how to track a Russian satellite. Fred Whipple explained delays in development of the cameras for optical tracking. It was during this meeting that the delays in delivery of the cameras prompted Richard Porter to say, “I have a number of times threatened to go up to Stanford and beat on tables. … Fred [Whipple] has so far frankly discouraged my doing so.”

At the thirteenth meeting of the TPESP, on October 22, 1957, it was reported that delivery of optics from Perkin Elmer had been increased and brought forward.

The fifteenth meeting of the TPESP, on January 7, 1958, demonstrates the poverty of information about the Sputnik s’ orbits. Whipple said, “We’ve not had a scrap of radio information.” Richard Porter, who headed the panel, said, “We may have underestimated again the difficulty of tracking and photography.” Pickering said, “The Soviet thing caught everyone off base” (page 34).

That the Soviets were also conducting the same basic science experiments and were interested in ionospheric refraction, tracking, and propagation effects comes from Selected Translations from Soviet-Bloc International Geo­physical Year Literature. Artificial Earth Satellite Observations (New York, U. S. Joint Publication Research Services, 1959) and Selected Reports Presented by the USSR at the Fifth Meeting of the Special Committee for the International Geophysical Year (New York, U. S. Joint Publication Research Services, 1958).

Details of the optical tracking program can be found in the annual reports of the SAO for 1961 and 1963.

Green and Lomask (Vanguard—A History, NASA History series SP4202) describe John Mengel’s actions when Sputnik was launched (page 35).

Chapter eighteen: Telstar

Documents drawn on for the launch of Telstar:

Satellite ground tracking station, Andover, Maine, Engineering notes: Tel­star July 9, 10, and 11, 1962. The document gives details of the count­down (page 188), for example, loss of calibration by the ground tracker at 1220 UT, power supply trouble at 2317 in the upper room of the com­munication antenna, etc. … (box 85080302 – AT&T archives).

Memorandum for the Record from John Pierce, Rudy Kompfner, and Chaplin Cutler on Research Toward Satellite Communication, and Research toward Satellite Communication (page 189). Both are dated Jan­uary 6, 1959, and deal with a research program directed in general at acquiring the basic knowledge for satellite communication by any means and specifically at aspects of passive Echo-type satellites. A fuller version of the research memo was written on January 9, 1959 (AT&T archives).

In this chapter references to what NASA officials said or did (pages 191 to 198) comes from documents in the NASA History Office or George Washington University. These were shared with me by David Whalen.

They include:

Memorandum for the Record, October 31, 1960, by Robert G. Nunn, special assistant to the administrator. This summarizes a meeting between NASA officials and James Fisk, president of Bell Telephone laboratories, and George Best, vice president of AT&T The purpose was to discuss Bell’s plans for “Transoceanic Communication via satellite.” It opened with T. Keith Glennan, NASA’s administrator, saying that Bell had not considered the “facts of life” with respect to vehicle availability. The meeting discussed policy issues in some depth, including finance and whether or why public money should be spent on communication satellites.

Memorandum for program directors, February 24, 1961. Subject: Guide­lines for preparation of preliminary Fiscal Year 1963 budget. On the sub­ject of communication satellites, it said to assume no funding of opera­tional systems; adequate provision should be made for back-up vehicles; and no development of passive communication systems.

Minutes of the administrator’s staff meeting: November 30, 1960; Decem­ber 1, 1960; December 8, 1960; January 18, 1961; January 26, 1961; Feb­ruary 2, 1961; March 2, 1961; May 25, 1961; June 1, 1961; June 12, 1961; June 15, 1961;June 22, 1961;June 29, 1961.

Technical details about Telstar and the attitudes and opinions of the Bell engineers were gleaned from the following:

“Project Telstar, Preliminary report Telstar 1 July—September 1962.” (AT&T archives).

Telstar—The Management Story, by A. C. Dickieson (unpublished manu­script, July 1970). Dickieson was the project manager for Telstar (AT&T archives).

Extracts from a manuscript by D. F. Floth. Chapter on Telstar Planning: January-May 1960 (AT&T Archives 84-0902).

Each quotes extensively from memos that the writers had access to.

The discussions of technology in the chapter come from a mixture of sources, including documents in the Hughes Aircraft Company’s archives.

Helpful textbooks include:

Satellite Communication Systems Engineering, by Wilbur L. Pritchard, Henri G. Suyerhoud, and Robert A. Nelson (Prentice Hall, 1993).

The Communication Satellite, by Mark Williamson (Adam Hilger, 1990).

Move Over, Sputnik

It was pretty tense because we knew that everybody was watching us, not only this country, but really the whole world, because here the Russians were making a big propaganda hit of how they were launching satellites and we were dropping rock­ets in a ball of fire on our launching pad. We did launch success­fully, at the end of January. That was a very interesting period to live through.

—William Pickering from a transcript of an oral history in the archives of

the California Institute ofTechnology.

I

n the late 1950s, there was no meeting of minds across the ideological divide.

“The country that gets a manned satellite into space first will be the undisputed master of the entire world. At the present time there is no defense against such a weapon. A satellite in a two-hour pole-to-pole orbit will pass over every part of the world every 24 hours [actually every 12 hours], and the launching of a guided missile against our cities would be a simple matter. Who is to control outer space? Russia? Or the United States?”

So wrote the editor of the Phoenix Republic sometime between Presi­dent Eisenhower’s announcement that the United States would launch satellites and Sputnik Vs arrival in orbit. Though more alarmist than many, the editor expressed a not uncommon fear.

Probably that fear was fueled by extracts from Soviet publications that appeared in the American press, such as the following from Soviet Fleet, a naval paper.

“American imperialists and their henchmen dream of using the pos­sibility of creating an artificial satellite… to set up outer world bases from which it would be possible to deliver attacks against countries of the democratic camp, and to hit the selected objectives.”

Amidst such rhetoric as well as the more measured and weighty crit­icisms of the New York Times, Sputnik II was launched on November 3, 1957. It was the second of three blows that year to America’s perception of its technological supremacy. The third, a month after the second Soviet satellite, would be self-inflicted. Sputnik II prompted yet more questions in

Congress, more headlines, more soul-searching editorials. Congressional critics urged Eisenhower to appoint a missile czar and pour money into education. On the Monday after the launch, Senator Lyndon Johnson spent the day closeted at the Pentagon. On Thursday, Eisenhower went on national television, attempting to reassure Americans that the country was secure. He emphasized the strategic importance of the Air Force, telling his audience that the United States Air Force was as effective as missiles.[8]

But the event that presaged America’s entry to the space age came on Friday, November 8, when Neil McElroy, the secretary of defense, directed the Army to prepare for a satellite launch as part of the Interna­tional Geophysical Year. The Vanguard team, however, was to get the first shot.

That shot took place on December 6. The countdown went smoothly; the launch was a disaster, one that was felt all the more keenly because, unlike the Soviet launches, it took place in full view of the world. Before the entire world, the rocket lifted about two feet off the ground and then burst into flames fourteen stories high. The explosion threw the third stage and satellite clear. The satellite landed on the beach. Its bent antenna beeped to a stunned audience.

J. Paul Walsh, the deputy director for the Vanguard project, had relayed the news over the telephone to listeners at the Naval Research Laboratory. His account was succinct: “Zero, fire, first ignition—explo­sion.”

The moment the news reached New York, there was a dash to unload Martin stock and that of other aerospace companies (though Lock­heed gained). At 11:50, the governors of the stock exchange suspended trading. The next day’s headlines in London included the ignominious words Flopnik and Kaputnik. Humor bolstered America, and people ordered Sputnik cocktails: one part vodka, two parts sour grapes.

There were to be worse failures. Astronauts and cosmonauts would die. In such a complex, unknown, and risky undertaking such disaster was (and is) inevitable. But this one had to hurt. The space community gritted its teeth and prepared for another launch. On January 27, 1958, Vanguard came within fourteen seconds of launch. The attempt was aborted. There was a problem with the second stage. Now, though, America had only four days to wait.

Immediately after McElroy’s direction of November 8, General Medaris, who headed the Army Ballistic Missile Agency (ABMA) in Huntsville, Alabama, had called Pickering, Homer Joe Stewart, and others to a meeting with himself and von Braun. The question was how to carve up the work to achieve a launch sometime toward the end of January, 1958.

Medaris assigned responsibility for the first stage of the launch vehi­cle to von Braun. This was the Redstone rocket, a redesigned and more powerful version of the V2. JPL was given responsibility for the satellite, the tracking stations, and the three upper stages, which would be solid – propellant rockets that the lab had developed. The entire launch vehicle was called Jupiter C.

JPL already had the tracking stations and the rockets because of its ongoing work with the Army exploring designs that would allow a missile to reenter the atmosphere without burning up. But they needed a satellite.

Pickering turned to Van Allen. They had previously talked infor­mally about whether Van Allen’s payload could be modified for an Army launch. Independently, Van Allen had talked in 1956 with staff at ABMA about an alternative, should Vanguard not be ready in time for the I GY. Van Allen would have known that delay was a possibility because Van­guard’s technical director, Milton Rosen, had briefed the IGY’s satellite panel about the technical difficulties with the rocket.

Therefore, once McElroy gave the army the go-ahead, Pickering sought permission from the IGY and Van Allen to prepare Van Allen’s payload for an Army launch. The IGY was the easy part. It was more diffi­cult to reach Van Allen, who was on a research vessel in the Antarctic. There Van Allen wrote in his notebook that Sputnik was a “brilliant achievement.’’ His reaction (and Guier’s) was in contrast to Rear Admiral Rawson Bennett’s comment to NBC that Sputnik I was “a hunk of iron that anyone could launch.”

On the deck of his cold, distant ship, Van Allen felt out of touch with the review of the U. S. program that was taking place. He was concerned that his group might miss a launch opportunity.

Van Allen was particularly worried when JPL acquired responsibility for the satellites (which became known as Explorer), fearing that the lab, which he perceived as very aggressive, would try to take over his experi­ment. His consolation was the confidence he had in Bill Pickering.

Pickering, in fact, went to considerable trouble to contact Van Allen, first with messages via the Navy. When that didn’t work, Pickering recalls that someone suggested Western Union. That succeeded. Van Allen cabled his agreement that his payload should be modified for an Army launch. His assistant, George Ludwig, picked up the bits and pieces around the lab­oratory and, in Pickering’s words, hightailed it out to Pasadena and the Jet Propulsion Laboratory.

The launch was scheduled for the end of January. This time, there was no formal prior announcement, though there were plenty of leaks. Journalists were on the alert. Shortly before the launch, Pickering’s staff were telling callers that Pickering was in New York. A wire reporter, keen to be sure that Pickering was where he was said to be and not in Washing­ton D. C. or at Cape Canaveral, turned up at Pickering’s New York hotel. “Just checking,” the reporter told him.

Days later, Pickering was in Washington—at the Pentagon with von Braun, Van Allen, senior Army personnel, and the secretary of the Army. They were waiting for the launch attempt of what would become 1958 alpha /, better known today as Explorer I. High winds had delayed the shot for twenty-four-hours. But on the evening of January 31, it seemed likely that it would go ahead.

Periodically someone would call the Cape to see how the count­down was faring. At T minus 45 minutes, launch controllers halted the countdown because engineers thought there was a fuel leak. After eigh­teen minutes, they decided there had been a spill during fueling and wiped up the mess. The countdown resumed. The servicing structure rolled back, and its lights went out. Now a search light picked out the silver-gray missile as a Klaxon sounded in warning.

At 10:48, Jupiter C lifted off. When the Redstone finished its burn, von Braun said to Pickering, “Well, now it’s your bird.” The bird had apparently been injected safely into orbit. But to be sure, they had to wait. Not before the tracking station in California picked up the satellite’s signal would they be confident that the spacecraft was orbiting. The only other people on the planet who could really know what that wait was like were in Kazakhstan.

Frank Goddard, at the Goddard Space Flight Center, was waiting to hear from California. Pickering kept a phone line open to him. They had predicted when they should hear the satellite. They waited as the minutes ticked past the time when they should have heard its signal. Pickering felt the glares on his back. Eight minutes after their predicted time, California confirmed that the satellite was in orbit, and the satellite completed its first orbit in the early hours of February 1.

Now von Braun, Van Allen, and Pickering were whisked through the rainy streets of Washington to a press conference at the National Academy of Sciences. They walked into a barrage of lights and microphones. “We didn’t know what we were getting into,” Pickering later recalled. “The place was jammed to the rafters. It was very exciting.”

The news was relayed to President Eisenhower in Augusta, Georgia, where he was enjoying a golfing vacation. He said, “How wonderful.”

Within minutes of the news reaching Huntsville, thousands of people took to the streets, honking car horns and carrying placards that read, “Move over Sputnik, our missiles never miss.”

America had entered the space age.

Before Sputnik I, the United States had planned that its first attempt to launch a full-scale satellite would be in the spring of 1958 and that four test vehicles carrying grapefruit-sized test satellites would be launched in the autumn of 1957. There were hopes that one would attain orbit. In the event, Vanguard put its first grapefruit in orbit on March 17, the day that Frank McClure called Guier and Weiffenbach into his office to discuss how their computational and statistical approach to tracking could serve navigation satellites.

By then, the space community was growing more comfortable with the techniques of satellite tracking. Yet during 1957 they had asked them­selves how they would track all the spacecraft if as many as six were to be launched each year. The question arose in 1957 as the satellite advocates tried to persuade their colleagues to endorse a continuing space program beyond 1958. Now, when TRW’s Space Log reports that by the end of 1987 there had been 2,979 known successful satellite launches (not includ­ing those deployed from the shuttles), that concern exemplifies the adage that the past is another country.

To today’s politically minded citizens, however, yesterday has familiar traces of home, namely budget battles, sniping between participants, and press relations.

By the beginning of 1956 the IGY’s total budget for the satellites and tracking was $19,262,000 an amount that approximately equaled its bud­get for everything else. This did not include the cost of developing and building the launch vehicles. Twelve satellites had been proposed by the scientists. The administration had announced ten in July 1955. By mid 1956, the scientists could count on six but, conservatively, were selecting only four for full development within the IGY’s timetable. All this took place within the context of Defense Department’s budget skirmishes and the rising costs of Vanguard.

The satellite panel was warned to keep the reduction confidential lest it damage America’s international prestige. Some clearly thought that this warning should not be heeded, because stories about the reduced pro­gram trickled out to the press, as did Fritz Zwicky’s comments to the American Rocket Society in November 1956 that “all kinds of jealousies, bureaucracies, and buck passing” were hindering the American satellite program. Many newspapers complained that the Navy should not have got the job of launching a satellite, and others reported on delays in placing of contracts for basic components.

Relations with the press were, in general, a contentious issue be­tween the IGY and the Department of Defense. The scientists grew increasingly irritated because, in their opinion, the Defense Department’s publicity machine made the project look like a military exercise. Not a dif­ficult job given that the Naval Research Laboratory was developing the launch vehicle and that some payloads were being prepared by scientists in defense laboratories. Nevertheless, and despite the fact that many of the university-based scientists had professional relationships at some time with the military, the scientists were determined to ensure that results of exper­iments were published in the open literature so that their international sci­entific relationships would not suffer. One can’t help but wonder what Soviet scientists were going through.

Amidst these concerns perhaps the most intriguing is the one that emerges in a flurry of correspondence in early 1957 that docu­ments that the IGY scientists feared that the Department of Defense would cancel the program once one satellite had been launched suc­cessfully

Nevertheless, planning for four satellites continued. And the space advocates succeeded in their campaign to convince their less enthusiastic colleagues to recommend that the satellite program continue after the IGY. As late as the day before the launch of Sputnik, this was not certain. But Sputnik, of course, changed everything for the space program. Like navigation, meteorology benefited.

Navigation section

Individuals interviewed for the navigation section are as follows:

Bob Danchik* (Transit’s penultimate project manager), Bill Guier* (physics), George WeifFenbach* (physics), Lee Pryor* (software develop­ment and Transit’s last project manager), Carl Bostrom* (physics and later the director of APL), Henry Elliott* (antennas), Lee Dubois* (command, control, and tracking), Charles Pollow* (assistant program manager), Lau­rence Rueger* (time and frequency systems), Tom Stansill* (receivers), Russ Bauer* (software), Charles Bitterli* (software), Harold Black* (physics/orbital mechanics), Ben Elder* (memory designer), Eugene Kylie* (receivers), Barry Oakes* (rf systems), Charles Owen* (mechanical design), Henry Riblet* (antenna design on Transit), Ed Westerfield* (receiver design), John O’Keefe (satellite geodesist), Gary Weir (naval his­torian), Commander William Craft (commander and director of seaman­ship at the U. S. Naval Academy, in Annapolis), Brad Parkinson (GPS pro­ject manager), Group Captain David Broughton (director of the Royal Institute of Navigation), and Dave Smith (satellite geodesist, currently at the Goddard Space Flight Center, Greenbelt, Maryland).

An asterisk denotes that an individual was a member of the Transit team.

Some of the above were interviewed in great depth and over many hours, weeks, and in the case of Guier and WeifFenbach, months; a very few spoke to me for as little as half an hour.

Chapter nineteen: The Whippersnapper

What Pat Hyland thought about Syncom’s early development is found in an extensive video interview recorded on December 14, 1989 (page 199-201). Copy available from НАС.

HAC’s early views on the commercial opportunities of space come from Bob Roney (page 201).

Frank Carver’s request that Harold Rosen look for new business ventures (page 201) was remembered by Harold Rosen and Bob Roney in their interviews with me.

The account of Rosen’s actions and discussion with Williams come from my interview with Rosen (page 202).

The account of Roney’s recruitment of Williams comes from my inter­view with Roney (page 202).

The account of Rosen’s efforts to tempt Williams back to Hughes comes from my interview with Rosen (page 202).

The technology in pages 203 and 204 is my distillation of the technical information in a number of memos, proposals, textbooks, and interviews.

Rosen s attraction to Southern Californian beach parties is his own recol­lection in an interview with me (page 204).

Sydney Metzgers comment (page 204) was made during an interview dated December 5, 1985, when he said, “When we (RCA) heard ofSyn – com we could have kicked ourselves for not thinking of a spinner at syn­chronous altitudes since RCA had the very early spinner experience.” Metzger, who worked for RCA, joined Comsat in June 1963 as the man­ager of engineering (НАС archives 1993-50 Box 1).

Comments on the TWT for HAC’s 24-hour satellite made in interviews with Tom Hudspeth, Rosen and Roney (pages 204 — 205).

The date that Leroy Tillotson sent his proposal for a medium-altitude satellite to Bell’s research department (page 205) is given in A. C. Dick – leson’s book (see notes for chapter 18).

Carver’s and Puckett’s immediate views ol Rosen’s proposal were given in Rosen’s interview with me (page 205).

A memo from A. S. Jerrems to F. R. Carver on September 17, 1959, reminds Carver of a meeting planned for September 23 to work up a pre­sentation on communication satellites for Allen Puckett (page 205) (НАС archives).

Rosen and Williams first describe their satellite in “Commercial Commu­nication Satellite,” October 1959, by H. A. Rosen and D. D. Williams (page 205), and in Preliminary design analysis of communication satellite, October 1959. This paper reviews the torque box design that Harold Rosen and Don Williams put forward for a 24-hour satellite in September (From НАС archives).

Sam Lutz examined the Rosen Williams idea. His evaluation appears in a memo from S. G. Lutz to A. V. Haeff, October 1, 1959, Evaluation of H.

A. Rosen’s commercial satellite communication proposal (From Bob Roney) (page 207).

A memo from A. S. Jerrems of October 9, 1959, confirms the establishment of a two-week-long intensive study of the Rosen proposal (page 207).

A memo from S. G. Lutz to A. V. Haeff on October 13, 1959. Subject: Eco­nomic aspects of satellite communication gives Lutz’s opinions (page 207).

Memo from J. H. Striebel to A. V. HaefF of October 22, 1959. Subject: market study for a worldwide communication system for commercial use shows more of the thinking at НАС (page 207).

Lutz’s second evaluation of the Rosen Williams proposal appears in a memo from S. G. Lutz to A. V. HaefF of October 22, 1959 (page 207). Subject: commercial satellite communication project; preliminary report on study task force.

A memo from L. A. Hyland to A. E. HaefF and C. G. Murphy of October 26, 1959. Subject: communication satellite orders an immediate and com­prehensive study should be made of patentable potentialities and NASA’s position should be ascertained (page 207). A number of subsequent memos show that Hyland’s instructions were carried out. Invention dis­closure was November 2, 1959.

A memo from D. D. Williams to D. E Doody on November 23, 1959 described Williams’s talks on November 5 with Homer Stewart, then at NASA, during which Williams emphasized that Hughes wished to main­tain its proprietary and patent rights and the company’s desire that the project should be undertaken as a commercial venture. The two also dis­cussed technical issues (page 208).

An interesting aside given the later legal action over patents between НАС and NASA is found in a memo from David Doody to Noel Hammond say­ing that should a 30-day analysis then being undertaken by the company show the 24-hour satellite to be feasible, Hughes would attempt to win a contract from NASA and would proceed with filing a patent application prior to contracting with NASA. He said further that the company would not yet enter the communication field or approach communication compa­nies with the proposal. He further wrote, “We will take our chances on retaining title to the inventions that have been made to date, but should NASA insist on taking title as a result of supporting the development, the company wifi go along with NASA since it does not intend to use resulting patents primarily for the purpose of enhancing its patent holdings.” This view is at odds with the decades-long battle that Hughes fought with NASA.

In September 1959, a barrage of technical memos begins covering topics such as dynamic aspects of communication project, feasibility investigation

of payload electronics. The technical memos mushroom during the fol­lowing years.

Despite Hyland’s decision not to commit funds to the 24-hour satellite (page 208) Rosen and Williams write “Commercial Communication Satel­lite”, January i960y by H. A. Rosen and D. D. Williams. By now the 24- hour satellite has the familiar cylindrical shape.

A memo from Robert Roney to A. E. Puckett of 27 January 1960. Sub­ject: communication satellite review analysis (From Bob Roney) describes yet another review of the Rosen/Williams idea (page 208).

On March 23, 1960, Williams wrote to Hyland, saying that he was pleased by Hyland’s decision to fund the commercial communication satellite. He wrote, “It is my understanding that the program will ultimately be financed by sources of capital external to the company. As one of the inventors of the system, I would like to invest in it myself if possible. I enclose a cashier’s cheque for $10,000. While I realize that this amount will not go very far, I think it can be multiplied by 100 if the company is willing to permit investment by its employees.” This was after Rosen, Hudspeth, and he decided to find some of their own money for the pro­ject (page 209).

A memo from Allen Puckett to D. E Doody dated March 7, 1960 details the requests by Williams, Rosen, and Hudspeth to be released from their usual patent agreements should Hughes not go ahead with the develop­ment of a communication satellite. Puckett states that their request is rea­sonable (page 210).

Details of Rosen’s attempts to raise money from various sources are from my interviews with Rosen.

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.