Category Dreams, Technology, and Scientific Discovery

The International Geophysical Year

T

he International Geophysical Year (IGY) was an epical scientific event. Through­out the community of geophysical researchers, it provided an infusion of funds and an integrating mechanism for hundreds of individual efforts that were made pos­sible by the rapidly evolving technologies of the time. The various coordination and sharing arrangements that were established for the IGY facilitated the interchange of data and information at a level far beyond that which would have occurred otherwise, and many of those arrangements have continued to the present day.

IGY inception and early planning

This IGY was not the first comparable international effort, although the scale of the new endeavor went far beyond that of previous ones. By the late nineteenth century, there had been a growing realization that the need for observations was global in nature, far beyond territorial boundaries, the intellectual resources of individuals, and even entire countries. This realization, coupled with the newly evolving technologies, was pushing the frontiers of human exploration over more and more of the Earth’s surface.

As a part of that expanding vision, two large-scale international efforts were mounted to apply the combined resources of the world’s leading scientists to exam­ine broad geophysical questions. The First International Polar Year, in 1882-1883, involved a collaboration of scientists to examine the geophysics of the Polar Regions, with concentration on the Arctic. That program included the establishment of a number of meteorological, magnetic, and auroral stations and their operation for about a year.

The Second International Polar Year was conducted during 1932-1933, the golden jubilee of the first. It repeated the earlier endeavor but extended its scope by adding ionospheric observations and by including a substantial component in the Antarctic.

67

OPENING SPACE RESEARCH

Подпись:Those two collaborations provided some of the inspiration, and served as a model, for the International Geophysical Year—1957-1958. That third endeavor, marking the silver jubilee of the Second Polar Year, took place during the period 1 July 1957 through 31 December 1958.

Much has been published over the years to document the planning and achievements of the IGY.1,2’3’4’5’6 Thus, only a brief summary is provided here.

An informal and largely spontaneous dinner party at the Van Allen home in 1950 served as the springboard for the IGY. Van Allen and his family were living at that time in a rented house at 1105 Meurilee Lane, Silver Spring, Maryland, located just off Dennis Avenue near its intersection with Sligo Creek Parkway. This was just a 10 minute drive from Van Allen’s laboratory in downtown Silver Spring.

In addition to hosts James and Abigail (Abbie) Van Allen, that dinner party included Lloyd V. Berkner, Sydney Chapman, J. Wally Joyce, S. Fred Singer, Merle A. Tuve, and E. Harry Vestine. The story of that momentous dinner is best told in Van Allen’s own words:

Vestine [Harry Vestine, who had originally urged Van Allen to make the electrojet search] was delighted with our equatorial electrojet results, as was [Sydney] Chapman who was visiting the United States in early April 1950. On April 5, they visited APL in order to learn about the results at first hand. Chapman expressed an interest in getting together with us and with Lloyd Berkner and Wally Joyce for further discussions. I immediately called my wife to confirm a previously tentative plan that she would have the group for dinner at our home. During the day, she cleaned the house, prepared a splendid dinner, and managed to feed our two young daughters and tuck them into bed as the guests arrived.

The occasion turned out to be one of the most felicitous and inspiring that I have ever experienced. Berkner was one of the leading experts on ionospheric physics and telecommu­nications at that time, had been a member of the scientific staff of the first Byrd Antarctic Expedition in 1928-1930, and had extensive experience in international cooperation in sci­ence while a member of the U. S. State Department. Joyce was a distinguished geomagnetician who had published the well-known Manual of Geophysical Prospecting with the Magnetome­ter in 1937 and was, as I recall, on the staff of the National Research Council at this time.

The dinner conversation ranged widely over geophysics and especially geomagnetism and ionospheric physics. Following dinner, as we were all sipping brandy in the living room, Berkner turned to Chapman and said, “Sydney, don’t you think that it is about time for another international polar year?” Chapman immediately embraced the suggestion, remarking that he had been thinking along the same lines himself. The conversation was then directed to the scope of the enterprise and to practical considerations of how to contact leading individuals in a wide range of international organizations in order to enlist their support. The year 1957-1958, the 25th anniversary of the second polar year and one of anticipated maximum solar activity, was selected. By the close of the evening Chapman, Berkner, and Joyce had agreed on the strategy for proceeding.7

Van Allen’s wife Abbie had a slightly different recollection of the event. In a conversation with Tom Krimigis in July 2007 she related, “… she was talking with Chapman first thing after the dinner and he was waxing eloquently about the Inter­national Polar Year of 1932-33. She suggested, “isn’t it about time to have another one?” to which he responded, “Well, maybe we should.”8

CHAPTER 3 • THE INTERNATIONAL GEOPHYSICAL YEAR 69

Planning for this third international endeavor progressed steadily. In May 1950, some 20 scientists, including Chapman (who had by that time left Oxford University for the University of Alaska), further discussed the suggestion at a meeting at the Naval Ordnance Test Station (now Naval Air Weapons Station) near Inyokern, China Lake, California. Further discussions and conceptual planning occurred soon after that at a Conference on the Physics of the Ionosphere, hosted by the Ionospheric Laboratory of Pennsylvania State University.

A formal proposal based on those discussions was conveyed by Berkner and Chapman in July 1950 to the Joint Commission on the Ionosphere, an organization of the International Council of Scientific Unions (ICSU). During the rest of 1950,

1951, and much of 1952, the proposal wended its way through the various ICSU member organizations. In the process, the World Meteorological Organization was added to the list of supporting institutions. The program’s scope was expanded to include synoptic observations of geophysical phenomena over the whole surface of the Earth. At the Amsterdam meeting of the ICSU General Assembly in October

1952, its name was changed to the International Geophysical Year—1957-1958. It soon came to be referred to simply as the IGY.

Sydney Chapman best summarized the scope of the IGY, as finally conceived, in his general foreword to the first volume of the Annals of the IGY:

The main aim is to learn more about the fluid envelope of our planet—the atmosphere and oceans—over all the Earth and at all heights and depths. The atmosphere, especially at its upper levels, is much affected by disturbances on the Sun; hence, this also will be observed more closely and continuously than hitherto. Weather, the ionosphere, the Earth’s magnetism, the polar lights, cosmic rays, glaciers all over the world, the size, and form of the Earth, natural and man-made radioactivity in the air and the seas, earthquake waves in remote places, will be among the subjects studied. These researches demand widespread simultaneous observation.9

The Joint Commission on the Ionosphere recommended that a committee be es­tablished as a focus for the detailed planning. That committee was formally consti­tuted, also at the October 1952 ICSU meeting, as the Comite Speciale de l’Annee Geophysique Internationale (referred to universally as the CSAGI). The CSAGI cen­tral direction was entrusted to a bureau consisting of Sydney Chapman (President), Lloyd Berkner (Vice President), Marcel Nicolet (Secretary-General), Vladimir V Beloussov (Member), and Jean Coulomb (Member).

Four major meetings of the CSAGI, including representation from all participating nations, were held during the period 1953 through 1956 to coordinate overall planning by the many suborganizations and among the national programs. Those pivotal meet­ings tookplace at Brussels on 30 June-3 July 1953, Rome on 30 September-4 October 1954, Brussels on 8-14 September 1955, and Barcelona on 10-15 September 1956.10

During that early planning period, a series of Antarctic, Arctic, Regional, and Discipline Conferences were also held to serve as forums for integrating various

OPENING SPACE RESEARCH

Подпись:components of the program. The Antarctic and Arctic conferences concentrated on the overall programs for those two regions, whereas the Regional conferences addressed the Western Hemisphere, Eastern Europe, Eurasia, Africa (south of the Sahara), and the Western Pacific.

The CSAGI established 14 Discipline Groups to plan many of the details.11

Responding to the invitation from the first CSAGI conference in July 1953 to countries of the world to join in the endeavor, the U. S. National Academy of Sciences, acting through its National Research Council, quickly established a U. S. National Committee for the IGY, with Joseph Kaplan as its chairman and Hugh Odishaw as its executive director. That committee served as the focal point throughout the duration of the program for all U. S. IGY planning and operational efforts.

Many other countries responded quickly to the invitation and set up their own mechanisms for coordinating their internal contributions. A key player, the Union of Soviet Socialist Republics (USSR), was slow in reporting its commitment to the IGY, but by the spring of 1955, it had also done so.

Sputnik!

T

he announcement of the launch of Sputnik 1 by the Soviets (Figure 6.1), like a bolt of lightning, instantly changed the complexion of the space program. In many ways, it changed from a collaborative international scientific research program to a race to demonstrate technical superiority during the cold war. It is certainly true that scientists successfully maintained a strong scientific content throughout the early history of the Space Age. But the surprising demonstration of the superiority of Soviet rocket capabilities alarmed everyone in the United States, from military planners to the general public, and precipitated a huge effort to catch up.

Early indications of Soviet intentions

The Sputnik 1 launch should not have come as such a surprise. There were many hints before October 1957 that the Soviet Sputnik would be appearing in our sky. Many of those were missed, ignored, or downplayed by all but the most astute observers. With the supreme confidence of U. S. satellite planners, accompanied by growing public interest resulting from the open publicity surrounding the Vanguard program, and with the Soviet program lurking behind its curtain of secrecy, we were blinded to the growing possibility that the Soviets might be able to enter space before us.1

A substantial body of USSR books, scientific and technical papers, newspaper and journal articles, and items in the popular science magazines carried such hints between the early 1950s and 1957. They provided ample evidence of a growing interest in the subject within the Soviet political, scientific, and public arenas. They covered a wide range of topics, ranging all the way from fanciful dreams and speculations to highly detailed descriptions of equipment, studies of weightlessness, and development of

159

Подпись: 160

Подпись: FIGURE 6.1 Model of the Soviet Sputnik 1, launched on 4 October 1957. It consisted of a sphere about 22.5 inches in diameter, with four radio antenna rods, shown here in the extended position. (Courtesy of the NASA National Space Science Data Center.)

OPENING SPACE RESEARCH

flight procedures for cosmic exploration. Many of those articles included interviews with widely respected scientists and engineers.

There were other specific indications of a Soviet intent to enter space. As early as November 1953, A. N. Nesmeyanov, then president of the USSR Academy of Sci­ences, told the World Peace Council in Vienna, “Science has reached a state when it is feasible to send a stratoplane to the Moon, to create an artificial satellite of the Earth.”2

On 16 April 1955, six months after the International Union of Geodesy and Geo­physics passed its resolution that a satellite program be added to the International Geophysical Year (IGY) program, the Moscow newspaper Vechernyaya Moskva re­ported that the Soviet Union planned to launch such a spacecraft. The article went on to tell of concrete actions toward that end, including the creation of an Interde­partmental Commission on Interplanetary Communication, chaired by Academician Leonid I. Sedov and reporting to the Academy of Sciences. Membership on that com­mission included a number of preeminent Soviet scientists. One of the commission’s explicitly named tasks was to launch a scientific Earth satellite to study the effects of weightlessness and of ultraviolet and X-rays from the Sun and stars and to observe ice floes and clouds. Moscow radio reported that a team of scientists had been formed to build the satellite.

On 30 July 1955, the day following President Eisenhower’s announcement of the U. S. plan to launch a satellite as part of its contribution to the IGY, the Kremlin tentatively revealed that the USSR planned to launch satellites during the IGY.

CHAPTER 6 • SPUTNIK! 161

Sedov, heading the Soviet delegation to the Sixth Congress of the International Astronautical Federation in Copenhagen in October 1955, repeated the idea at a press conference that they, too, might be in the satellite game. Choosing his words carefully, he said:

From a technical point of view, it is possible to create a satellite of larger dimensions than that reported in the newspapers, which we had the opportunity of scanning today. The realization of the Soviet project can be expected in the comparatively near future. I won’t take it upon myself to name the date more precisely.3

Despite the press articles and Sedov’s statement, the USSR made no official de­cision on launching an IGY satellite until early 1956. It was then that the USSR Academy of Sciences quietly made an internal commitment to launch an IGY satel­lite within two years, and authorization for its development was made in the form of a decree of the Soviet Council of Ministers on 30 January 1956.

It turned out that that decision was not immediately followed by specific actions to meet the challenge. As the year progressed, progress was lagging, in the opinion of the chief USSR rocket designer, Sergei Pavlovich Korolev, and his allies. One of those associates, Mstislav Keldysh, appealed to the Soviet Academy of Sciences on 14 September 1956, only a year before the Sputnik 1 launch, for more vigorous action. During his presentation, Keldysh stated that they were considering “placing a live organism in the satellite—a dog.” He proceeded beyond that to tantalize them with visions of a flight to the Moon to take pictures of its dark side. His final remark to them was to urge, “It would be good if the Presidium were to turn the serious attention of all its institutions to the necessity of doing this work on time. . . . We all want our satellite to fly earlier than the Americans.”4 In spite of his appeal, the USSR satellite development continued to lag.

There continued to be external signs of the Soviet commitment. In September 1956, Ivan P. Bardin very clearly advised the attendees of the Fourth IGY COSPAR meeting in Barcelona, Spain, that the Soviet Union “intends to launch a satellite by means of which measurements of atmospheric pressure and temperature, as well as observations of cosmic rays, micrometeorites, the geomagnetic field and solar radiation will be conducted. The preparations for launching the satellite are presently being made.”5

By November, the Soviet plans were finally taking tangible form as a duo of satellite designs. Their pride and joy, the one intended to be launched first as their primary contribution for the IGY, weighed in at almost a ton and a half. It was loaded with an assortment of scientific instruments, including, among others, a magnetome­ter, photomultipliers, a mass spectrometer, ion traps, and photon and cosmic ray detectors.

OPENING SPACE RESEARCH

Подпись: 162A second satellite would be very simple—a 184 pound sphere carrying a pair of radio transmitters radiating at frequencies that would make it easy for radio amateurs and anyone else with a simple receiver to receive its signals.

Because of the slowness in getting the program up to full speed, and difficulties in getting all of the flight hardware to operate properly, the launching order was ultimately reversed. The second design became the first Sputnik. A new satellite design with the dog Laika was hurriedly prepared and launched as Sputnik 2 on 3 November 1957. The fully instrumented satellite, after a first launch failure on 27 April 1958, was finally launched on 15 May 1958 as Sputnik 3.

The decision to choose the simple 184 pound satellite for first launch was made only because of the Soviets’ intense desire to beat the United States into space. In a much later interview, Gyorgi Grechko, an engineer who worked on the early USSR satellite program, stated:

But, these devices [referring to the instruments for what became Sputnik 3] were not reliable enough, so the scientists who created them asked us to delay the launch month by month.

We thought that if we postponed and postponed we would be second to the U. S. in the space race, so we made the simplest satellite, called just that—Prosteishiy Sputnik, or “PS.” We made it in one month, with only one reason, to be first in space.6

It appears that the Soviet decision to launch the simple version had already been made by the time of a meeting at the U. S. National Academy of Sciences in June 1957. At that meeting, the Soviet Bardin provided a document to Lloyd Berkner, the Comite Speciale de l’Annee Geophysique Internationale (CSAGI) reporter on rockets and satellites, titled USSR Rocket and Earth Satellite Program for the IGY. In the informal discussions outside the meetings, the Soviets talked quite openly of their plans, contrary to later popular claims.7

Within a month of that meeting, Radio, a Russian amateur radio magazine, included two articles giving a reasonably comprehensive description of that satellite’s intended orbit. It went on to tell how its approach could be predicted by receiving stations anywhere along its path, and how its 20 and 40 MHz signals could be received. They went so far as to include instructions for building shortwave radio receivers to pick up the signals, plus a direction-finding attachment for locating the satellite.

Those articles appear not to have entered the consciousness of U. S. scientists and officials until they were introduced by the Soviet attendees at the CSAGI Conference on Rockets and Satellites during the week of the first Sputnik launch. The primary response to that information at the conference was shock and irritation that the Soviets had departed from the agreed-upon frequencies for transmission. It was not interpreted as an indication that the launch was hard upon us.

Development of a suitable launching rocket was the key to orbiting an Earth satellite. The U. S. planners decided upon the development of a new launch vehicle derived

CHAPTER 6 • SPUTNIK! 163

from technology developed for nonmilitary scientific research. There were a number of reasons for that. One was the desire to keep the U. S. satellite program distanced as much as possible from any signs of military involvement. The more complete story of that decision is related in Chapter 7.

The Soviets had no such compunction. In those days, practically no significant high-technology activity was conducted within the Soviet Union that did not have either military or propaganda implications. They were nearing completion of the development of their first Intercontinental Ballistic Missile (ICBM), the R-7 (SS-6 Sapwood), by a team under the leadership of Sergei Korolev. On 15 May 1957, the first R-7 test launch resulted in an explosion upon ignition. Four additional attempts during the summer also failed. Finally, on 21 August 1957, an R-7 rocket flew 4000 miles over Siberia, landing in the Pacific Ocean near the Kamchatka Peninsula. After a second successful launch, Pravda on 27 August announced to the world that successful tests of an ICBM had been carried out.

It was after that success that Soviet premier Nikita Khrushchev finally yielded to arguments by Korolev and others and gave his approval for the launch of the satellite that had been under development for such a long time. On the centennial of Tsiolkovsky’s birth, 17 September, the Soviet government promised the world that a satellite would soon be launched.

Although the significance of the Pravda announcement in terms of the Soviet Union’s ability to deliver nuclear weapons to any point on the Earth was clearly rec­ognized, there was no expectation in the West that that rocket’s first major assignment would be to loft an instrumented payload into Earth orbit only a few weeks later. U. S. scientists and engineers remained supremely confident that they would launch a satellite into Earth orbit long before the Soviet promise materialized.

A call from the Jet Propulsion Laboratory

In addition, the satellite design had to be completed and built. For the 18 days imme­diately following the Sputnik launch, I was outside the circle of frenzied discussions, negotiations, and design work mentioned above. That situation changed dramatically on Tuesday, 22 October, when I received a telephone call from JPL’s Eberhardt Rechtin. Following up on our discussion in May of matching an abbreviated version of our cosmic ray package to their Microlock system, he briefly outlined their evolving plans and arranged to visit us the following day.

The JPL visitors joined us in Van Allen’s conference room at 4:00 on Wednesday afternoon. Eb happened to be too ill to travel on that day, so Henry Richter came in his place, accompanied by Walter (Walt) J. Downhower. The Iowa attendees at the beginning of the meeting were Ernie Ray, Frank McDonald, Kinsey Anderson, and me (Van Allen was in the South Pacific). Robert (Bob) Parent, from Vern Suomi’s group at the University of Wisconsin, was on campus for a different meeting that morning, and he joined us.

To open the meeting, Henry Richter stated that Eb Rechtin would be visiting the secretary of the Department of Defense two days hence to offer their services, and they wanted to be able to present specific information on scientific instruments to be carried. After that brief introduction, the other Iowa attendees left, and the JPL visitors and I got down to business. They outlined the overall concept of their satellite and the physical environment to be endured by the instruments. We developed a plan for the integration of our instrument that would meet the combined objectives of ABMA, JPL, and the University of Iowa. Their key objective was to place a useful satellite in orbit as quickly as possible. Our primary objective at Iowa remained as before—to conduct the geographically broad survey of cosmic ray intensity that Van Allen had originally proposed.

The discussion quickly led to an understanding that two packages would be launched. The abbreviated package for an initial rush effort would include the basic cosmic ray instrument that we had been discussing at ABMA. That would be followed

OPENING SPACE RESEARCH

Подпись:by a second launch of our complete cosmic ray package, essentially as it had been designed for Vanguard.

Several other instruments were being considered for inclusion—thermistors to record temperatures at several places in the satellite and an abbreviated version of Vern Suomi’s solar radiation balance instrument from the University of Wisconsin. Following the meeting, however, after Parent had returned to Madison and briefed Suomi, Suomi stated that it would be difficult or impossible to achieve their scientific objectives with an abbreviated package. Since their full instrument with its internal storage could not be accommodated on either flight, further consideration of their experiment was dropped in favor of relatively simple micrometeorite instruments being developed for Vanguard by Edward (Ed) Manring and Maurice Dubin at the Air Force Cambridge Research Center (AFCRC) near Boston.

It was only later that I realized that I had been drawn, unknowingly, into a major institutional power play. As detailed earlier, we had been working since early in 1957 with Ernst Stuhlinger, Josef Boehm, and their staffs at ABMA on their design for a Jupiter C satellite. Although those discussions included JPL engineers, our understanding was that their participation was for the sole purpose of integrating their Microlock system into the ABMA satellite. It remained our full expectation, as well as that of ABMA’s von Braun, Stuhlinger, Boehm, and their staffs, even until early November, that ABMA would have the lead in developing the satellite, including integrating our instrument.

But Pickering and his staff at JPL had their eyes set upon the satellite task. They believed that the payload logically fit within JPL’s mission, as their organization was shifting away from missile development toward high-technology electronics.

So it was the JPL engineers, instead of ABMA staff members, who visited us on 23 October. As my working relationship with JPL progressed over the next few days, I became increasingly troubled that the ABMA engineers were not participating in our discussions. I mentioned that to Pickering during a conversation on 5 November, but he assured me that his staff were working closely with von Braun’s people, and that it was fully appropriate that I continue to work with the JPL staff. I accepted his assurance.

As I learned later, at the time of that conversation, the ABMA people were com­pletely unaware of the JPL aspirations and separate satellite development effort, even though it had been under way for most of a year. They did not learn of that until 9 November 1957, the day after the directive was issued for the Army to proceed with the Jupiter C satellite program.

On that date, General Medaris held a meeting at Huntsville of ABMA, JPL, and Army personnel to clarify the assignment of roles and responsibilities. It was at that time that Pickering made his move. He asked Medaris for a brief private discussion

CHAPTER 8 • GO! JUPITER C, JUNO, AND DEAL I before the meeting, where he convinced the general that the satellite development role should be assigned to JPL.

Pickering went into that premeeting discussion armed with a number of argu­ments. The JPL work on communications, as exemplified by the Microlock devel­opment, was a strong factor. Furthermore, Pickering believed that they had earned the satellite role through their work with ABMA on Orbiter, Jupiter C, the RTV, and, especially, their efforts during the previous two years in helping to keep the Army satellite possibility alive. He must also have pointed out that greater efficiency and probability of success would result from having the entire upper stage endeavor, including the satellite, located at a single institution rather than to have the rocket cluster/satellite/tracking/telemetry interfaces spread between the two organizations.

Pickering undoubtedly revealed during that private premeeting discussion that JPL had been developing their own satellite and that they were much further along in its development than the ABMA staff. Their production of actual prototype hardware, complete except for the scientific instruments, attested to that fact. I suspect that the final arrow in Pickering’s quiver was the revelation that he had already arranged with me at the University of Iowa to shift our instrument from the Vanguard to the JPL satellite effort.

As author Clayton R. Koppes described the premeeting discussion in 1982:

Just prior to the meeting at which the roles would be assigned, Pickering asked Medaris for a few minutes alone. He argued that JPL should build the satellite. The general probably felt the laboratory could handle the electronics work better than Redstone, and he wanted to keep JPL actively in the Army’s orbit. Von Braun’s jaw dropped when Medaris and Pickering walked into the meeting and informed him of the decision, but the collaboration proved fruitful, and there was more than enough work for both teams. The quarter of an hour Pickering spent with Medaris was momentous. If Redstone had built the Explorer I satellite, it would have had a lock on both the missile [booster] and the satellite. JPL would have been relegated to a minor supporting role, chiefly in its [upper stages and] tracking network, from which it would have been highly unlikely to develop into a major space laboratory. Electronics, which had begun shouldering propulsion aside as the laboratory’s dominant activity during the Corporal weaponization, opened a window to space for JPL.8

In 1986, Pickering recalled that particular event in the following terms:

Medaris had a big meeting on 9 November with about twenty people, including Stewart, Froehlich, and myself. At that meeting, he announced that JPL would be involved [with the satellite design]. I think that came as quite a shock to the Germans. . . we could sense the reaction. But we worked very well together, right up to the launching.9

Additional accounts appeared. Author William E. Burrows described the event and its implications in 1998 in the following way:

Having gotten a green light on the launcher, von Braun and his colleagues assumed that they would get to develop the satellite as well. But they were wrong. Pickering, the chief of the Jet Propulsion Laboratory and a tough New Zealander who was educated at Caltech, believed

OPENING SPACE RESEARCH

Подпись: 220that given the lab’s involvement with Project Orbiter from the start, it had every right to continue participating in the push to reach space. He therefore convinced Medaris to award the satellite contract to JPL. The general undoubtedly believed that JPL, which by then had solid experience in both electronics and rocketry, could handle the assignment better than his own agency.

Whatever Medaris’ motive, though, his decision would prove to be a momentous one. By allowing JPL to design and build what would turn out to be America’s first spacecraft, he unknowingly played into the hands of a man who had already quietly decided that JPL needed to abandon work on tactical missiles and aim higher. Pickering wanted to build machines that would explore the solar system. As a university [administered] laboratory, he would explain years later, “It was quite clear to us that our future lay on the space program,” not in weapons work, which offered little intellectual challenge.10

I continue to be disturbed by the fact that I had been carried into JPL’s private satellite development effort well before ABMA was aware that JPL was seeking the satellite-building role. I believed that I was still operating within the ABMA-led collaborative satellite-building effort. If I had that period to live over, I would certainly call Stuhlinger to discuss the institutional ramifications. Although that would probably not have changed the outcome, it would have made the process more open.

Thus, in retrospect, during the period immediately following the Sputnik 1 launch, three possibilities existed for us at Iowa. First, we could have remained with the Vanguard program by rejecting the Army proposal altogether. Second, we could have continued with the ABMA-led planning for the satellite. Finally, we could have accepted the JPL offer. From the vantage point of later developments, the JPL alternative that we accepted was by far the best one. Had we remained with Vanguard, we would have had only one launch opportunity, as the small number of launch vehicles being procured for that program was fully subscribed for the vehicle-testing program and the launching of the scientific instruments that were being developed. As events unfolded, being second in the Vanguard experiment queue, our instrument would probably not have had a launch attempt until sometime in 1959. And history shows that the probability of a successful launch would have been small. The overall probability that we would have discovered the radiation belts ahead of the Soviets is essentially zero.

The second alternative would have been equally disappointing. The ABMA satellite plan included only a single satellite design—the highly truncated one. Although that might well have been successfully launched, we would still have had to depend on the Vanguard program for launching the full instrument with its onboard storage. That would very possibly not have succeeded, as speculated above.

The JPL promise of two launches, one quick one with the truncated instrument and a second one with our full instrument, was tremendously attractive. The decision to go with the JPL satellite program was felicitous—ABMA and JPL made good

CHAPTER 8 • GO! JUPITER C, JUNO, AND DEAL I 221

on their promise, including preparing and launching a third launch vehicle after the failed attempt to launch the complete instrument as Explorer II.

Once the decision was made for JPL to take the lead on the satellite, the arrangement worked well. Throughout, Von Braun insisted that the ABMA primary interest was in getting a satellite into orbit, rather than squabbling over roles. On one occasion, when he and some of his senior staff were flying to California in the ABMA Gulfstream for a meeting with the JPL staff, a discussion broke out about a particular question of roles. The staff wanted von Braun to insist on ABMA responsibility for an activity that JPL also wanted. Von Braun would have none of that. His response to them was to restate his general objective and to threaten the staff that, if they wanted to make an issue of the roles question, he would have the plane turn around and fly back to Huntsville. They continued on to Pasadena.11

To return to the chain of events, Richter and Downhower proceeded from our meeting in Iowa City to Washington, where they joined other JPL members on Thursday for a discussion with a bevy of Army and Navy officers at the Pentagon. The membership of that JPL contingent is a little unclear but probably included either Bill Pickering or Eb Rechtin, plus Jack Froehlich, Al Hibbs, Richter, and Downhower, and perhaps Homer Stewart. Rechtin met with Army Secretary Wilber M. Brucker and Defense Department missile coordinator William A. Holaday on Friday to present their Deal plan. On the following Monday, 28 October, Brucker tentatively agreed with the two-satellite Army/JPL approach. Late that afternoon, Rechtin informed me of that agreement, emphasizing that final approval would have to come from the secretary of defense.

It was during that conversation that Rechtin asked if I had the full authority to shift our experiment from Vanguard to the Army program. That question startled me. I had proceeded with full confidence that Van Allen would be enthusiastically in favor of the change if he had been present. After all, he had long viewed the Army’s Redstone rocket as an attractive launch vehicle. The two of us had made a substantial effort to design our Vanguard instrument package so that it could fit in either the Vanguard or the Jupiter C configuration. The satellite instrument that he and I had worked out with Ernst Stuhlinger and his people during the preceding spring and summer was identical in concept with that now being proposed for the first Army launch. And Rechtin had assured me that our full package would be flown on a second launch.

However, when faced directly with that query, I realized that I, a graduate student working on my master’s thesis on what was actually Van Allen’s proposal and project, did not really have the authority to make such a fundamental decision. I had to reply

OPENING SPACE RESEARCH

Подпись:to Eb that final approval for Iowa’s program change would have to come from Van Allen and offered to find a way to reach him.

The next morning, I contacted Mr. Sedwick in Commander Hearn’s office at the Naval Research Laboratory (NRL). His was the office responsible for coordinating the USS Glacier expedition to the South Pacific. He told me that unclassified mes­sages could be given for transmission to the nearest Naval Communication Center through Western Union, or to the commandant of the nearest Naval District. Clas­sified messages had to be hand carried to either the nearest Naval Communication Center or other military center. I called Eb Rechtin at JPL to convey that information at 11:00.

A little later in the day, Henry Richter called to tell me that the USS Glacier would not be arriving in New Zealand until 10 November and that they were attempting to reach Van Allen on shipboard.

During that discussion, Richter told me that they had received a copy of the letter carrying the army secretary’s approval of the program. With that, and anticipating Van Allen’s favorable response, I proceeded with my planning.

The following day, Wednesday, 30 October, Richard Porter, chairman of the U. S. Technical Panel on the Earth Satellite Program (TPESP), informed me that he had relayed the proposal for flying our cosmic ray instrument, plus the temperature­measuring and micrometeorite instruments, to the U. S. National Committee for the International Geophysical Year (IGY), and that the committee had approved. He further stated that he had discussed the change with Homer Newell, overall head of the Vanguard project at NRL, and that he had not raised any objection. Porter told me that a meeting of the TPESP was scheduled for 6 November and that he would relay our requirement for additional University of Iowa funding for the new work.

I was alarmed at one point during that conversation. Porter seemed to be under the impression that there would be only one flight if the first launch attempt succeeded. He heightened my concern by urging that we prepare our complete package for the first flight “if at all possible.” That was not the agreement I had with the JPL people. When I talked with Henry Richter the next morning, I voiced my concern about Porter’s comments, and reemphasized that the complete package with onboard storage was essential for at least one flight. Henry promised to check into the matter, and later reaffirmed the plan for the multiple launches.

Henry also told me during that conversation that JPL had received the first reply from Van Allen but that it was inconclusive. Because of the project’s secret classi­fication, many of the details could not be mentioned in unencrypted messages. To avoid the complications of handling classified messages, JPL had decided to obtain

CHAPTER 8 • GO! JUPITER C, JUNO, AND DEAL I 223

his approval through clear channels. The second terse message that Pickering sent to Van Allen on 30 October read simply:

To Dr. Van Allen. Would you approve transfer of your experiment to us with two copies in spring? Please advise immediately. Signed: Pickering.12

In the absence of information about the recent negotiations and decisions being made in the United States, even with his full knowledge of the earlier Jupiter C planning for our instrument, it was impossible for Van to understand what was being asked. He replied on 31 October:

Dr. W. H. Pickering. Unable to interpret your word transfer due ignorance recent devel­opments. Our apparatus for original vehicle nearly finished. Delighted prepare three sets non-storage type for JPL program. Signed: Van Allen.13

That message implies that Van Allen was still thinking in terms of the earlier work with the ABMA personnel, and that we would still be depending on Vanguard for the full instrument package. Pickering responded with a message that reached Van Allen on 2 November:

Present planning suggests [flying] most of existing equipment but transferring responsibility to us instead of carrying two programs. First experiment would be continuous. Second would be storage type. Suggest you phone me on arrival in New Zealand.14

The next day we all received another huge surprise—the launch of Sputnik 2 by the Soviets! Whereas we had been startled in October by the large weight and size of their Sputnik 1, we were dumbfounded by the size and sophistication of Sputnik 2. Our Vanguard satellites were expected to weigh only 23 pounds. Sputnik 1 had weighed 184 pounds. Sputnik 2 weighed an unbelievable 1121 pounds! Not only that, but it carried a live animal, the dog Laika, in addition to its array of scientific instruments. Even though it was revealed later that no provisions had been made for recovering the dog, and that she died of heat exhaustion after about seven hours in orbit, the feat represented a huge coup by the Soviets.

No longer could anyone assert that Sputnik 1 had been the lucky result of a short­term crash effort. It was obvious that the Soviets were very serious about the space business. The workers (and Khrushchev, immediately after the Sputnik 1 launch) clearly recognized the impact of what had now become a race to capture the attention of the whole world and were intent on demonstrating and exploiting their scientific and technological prowess. Even more profoundly, inclusion of the dog in Sputnik 2 suggested their strong interest in manned flight.

OPENING SPACE RESEARCH

Подпись: 224The pressures for an early American satellite mounted!

When the USS Glacier finally arrived in Port Lyttleton, New Zealand, on 10 Novem­ber, an additional message from Pickering awaited Van Allen:

Urgently need your approval on proposed change of Ludwig experiment. Porter committee has given their approval to proposal to change experiment to JPL and to modify experiments as agreed upon between Ludwig and JPL.15

Still somewhat uneasy about his lack of solid information, Van Allen addressed a commercial cable to me at Iowa City on 13 November, asking:

Question: Is Pickering plan for our experiment agreeable with you: Please cable answer IGY rep Christchurch.16

Ernie Ray immediately answered the cable on my behalf:

After high-level approval and obvious rearrangement of old program, George left town for extended stay. George quite happy Pickering plans. Hope you say yes.17

With that reassurance, Van Allen immediately wired Pickering:

Approve transfer our experiment accordance JPL plan.18

Thus, on 14 November, only six days after Department of Defense formal approval of the Army’s satellite plan, Van Allen provided his final approval, and we were free to put the plan into effect.

Van told me some time later that his hesitancy in agreeing to the telegram from JPL was to some extent due to his concern that they might be trying to take over the entire experiment, including the scientific analysis.19 Recently, he indi­cated that he was not as worried as he might have been, because of his belief in Pickering’s innate honesty. That confidence had developed during the many years that the two of them had worked together. Still, Van was much relieved to receive Ernie’s message. [2]

CHAPTER 8 • GO! JUPITER C, JUNO, AND DEAL I 225

Therefore, I accepted the JPL employment offer as the most expeditious way to proceed. During the rest of that frantic week, I had a number of conversations with Henry Richter, and with the JPL Personnel Office, to arrange for my employment. I was hired as an employee of the California Institute of Technology, working at JPL, with a title of research engineer, effective 18 November 1957, and at a starting salary of $700 per month.

Van and I have looked back on that decision on several occasions and agree in retrospect that it would have been much better if I had remained on the University of Iowa payroll. I would have been received at JPL more clearly as a senior colleague, rather than as a junior member of their own technical staff. As a practical matter, that turned out not to be a serious problem, as I could work around any issues by calling Van Allen and letting him work them out with Pickering. However, it did cause me considerable frustration, as the arrangement left me outside some of the project-related activities that would have been helpful.

Returning to the events of late October and early November, we used my pending employment at JPL as the official justification for a rush trip to Pasadena. Although it was carried on the books as a recruiting interview trip, its real purpose was for Deal project planning. In anticipation of Van Allen’s approval of the programmatic shift, I left Iowa City on 1 November, arriving in Pasadena late that evening. The next day, we worked out many of the details of the two payloads and of the ground system needed to track and communicate with them. JPL attendees at that meeting included, at various times, JPL director Pickering, project director Jack Froehlich, Eb Rechtin, and Henry Richter.

The primary result of that meeting was full agreement on the payload and ground station configurations for the two launches.20 The block diagram for Deal I appears here as Figure 8.2. With only minor changes, that configuration was the one eventually flown on Explorer I. A similar diagram was sketched at that time for the more complex Deal II, as described later.

Whereas the Vanguard system had included a single transmitter for tracking and data transmission, two transmitters were planned for both Deal I and Deal II to provide greater redundancy. As an added advantage, the system would employ both the Vanguard Minitrack ground network already being built by NRL, and the new JPL Microlock ground network, thus providing expanded data recovery and additional ground system redundancy.

We envisioned at the 2 November meeting that both the Deal I and II satellites would employ a low-powered transmitter radiating continuously at a frequency of 108.00 MHz, a power level of 0.01 to 0.02 watts, and with a linearly polarized asym­metrical dipole antenna. That system would be included primarily for Microlock tracking and telemetry reception, and secondarily for Minitrack interferometric

A call from the Jet Propulsion Laboratory
tracking. It would be expected to operate for two to three weeks. By the time Deal I was actually launched in January, the power had been fixed at 0.01 watt, and the design operating lifetime had been increased to “two to three months.”

A relatively high powered second transmitter was to be used on Deal I primarily for transmitting the cosmic ray data to the Minitrack ground network and for inter­ferometric tracking by those stations. It was to radiate continuously at a frequency of 108.03 MHz with a power level of 0.1 watt, for two to three weeks. It would use a circularly polarized X-shaped antenna similar in its mode of operation to the antenna that had been designed for Vanguard. Identical cosmic ray data streams were to be transmitted through both the low – and high-power systems.

The second Deal I experiment, conducted by Ed Manring and Maurice Dubin of the AFCRC, was to study micrometeorites. Micrometeorite density was of considerable scientific interest in its own right. Also, since the space environment was largely unknown, there was a desire to see if very small meteorites were dense and heavy enough to represent a hazard to this and future satellites.

The AFCRC instruments employed two detectors. The first, a microphone placed in spring contact with the satellite’s shell, was tuned to respond to micrometeorite impacts of 0.012 grams per centimeter per second and greater. In more precise terms,

CHAPTER 8 • GO! JUPITER C, JUNO, AND DEAL I 227

the minimum detectable particle mass would be 5.4 x 10-9 grams (assuming a particle velocity of 40 kilometers per second relative to the satellite). The output of that tuned amplifier, scaled by a factor of four, was transmitted via channel 3 of the high-power transmitter.

The second micrometeorite detector was a set of 12 parallel-connected grids of very fine wire. Each grid was wound on a flat bobbin so that it completely covered an area of about one square centimeter. If a micrometeorite of size five microns or larger impinged on one of those grids, its wire would be severed, causing an abrupt change in the resistance of the network. That resistance was telemetered via channel 3 of the low-power system.

The art and science of satellite design was in its infancy, and controlling the temperature of the instruments within the satellite shell represented unknown territory. There was a strong motive for verifying the accuracy of the many calculations and physical provisions for temperature control. To obtain that verification, measurements were made of the temperatures of the aft end of the main cylindrical satellite shell and of the transistor in the high-power transmitter (conveyed over channels 1 and 2 of the high-power system), and of the aft end and tip of the satellite’s front cone (over channels 1 and 2 of the low-power system).

At the 2 November meeting, we also tentatively agreed on plans for JPL to prepare three primary Microlock stations, one near the launch site at Patrick Air Force Base (Cape Canaveral, Florida) and two to be located at then unspecified locations, but “perhaps in Hawaii or the Philippines.” Use would be made of NRL’s full Minitrack stations in South Africa and Australia, plus somewhat simplified Minitrack stations in the United States, South America, New Zealand, Nigeria, and Morocco. That initially envisioned network configuration was eventually considerably modified, as described later.

At the end of that Saturday meeting, we agreed to give urgent priority to procuring certain components that might not be available from supplier’s stocks and might have to be manufactured. Other matters, including arrangements for the University of Iowa to provide the flight tape recorders, security, and financial arrangements, were also addressed.

I spent part of the next morning familiarizing myself with the Pasadena area so that I would be prepared to search for a temporary home when I arrived with my family. That Sunday afternoon, I returned to Iowa City.

The next day, I called Pickering with my proposed budget for the additional Iowa work. Our total budget for developing the satellite instrument as part of the Vanguard Project had been $106,375. I estimated that we would need an additional $39,100 if we were to modify the package at Iowa City. That included an additional $2000 for two

OPENING SPACE RESEARCH

Подпись:part-time employees for six months, $12,000 for additional equipment and supplies, $8000 for additional travel, and $12,000 for data reduction, which, we realized by then, had not been adequately funded in the original Vanguard budget. I also informed him that, if the work on preparing the instrument were shifted to JPL, the University of Iowa would need only the additional $12,000 for data reduction.

Those funding figures highlight an interesting aspect of the space program at that time. We, on our university campus, and JPL, in Pasadena, operated in entirely differ­ent financial realms. During the preceding seven years, the Iowa Physics Department under Van Allen’s leadership had developed a burgeoning space research laboratory from nearly nothing, on a shoestring budget. We were used to scrounging military surplus sources for resistors, capacitors, screws, and other parts, and to operating with a tiny staff of a few faculty members and graduate students and a handful of part­time students. Improvisation was the byword. The total University of Iowa budget of just over $100,000 for the Vanguard instrument (admittedly not including instrument shop expenses and overhead charges) was laughable by later standards, even within a university environment.

On the other hand, JPL was a major military contractor. By that time, it had a staff of over a thousand and was well versed in the business of developing and testing ultrareliable equipment to meet military standards. They were able to commit hundreds of people to the Deal project. A 20 November 1957 Deal organizational chart lists 26 people in senior supervisory and consultative roles, and about 100 others with major named responsibilities, split almost evenly between rocket preparation, satellite instrumentation, and ground stations and operations. Those individuals were backed by legions of technicians and other supporting staff.

This is by no means a criticism of JPL. We would have been terribly hard-pressed at Iowa to prepare our cosmic ray instrument in time to meet the Deal schedule. In addition, the JPL effort provided a much greater assurance of reliable performance for several reasons. The design of all elements of the configuration was worked over meticulously by the highly skilled and experienced JPL engineers, the testing was thorough, and the combined Microlock and Minitrack tracking and telemetry configuration was far more robust than the Vanguard Minitrack system would have been by itself.

In summary, the move of our instrument construction from Iowa City to Pasadena was the only viable arrangement under the circumstances.

On 6 November, immediately upon my return from Pasadena, my direct participation in the Deal project began in earnest with a long working telephone conversation with Henry Richter. He informed me that they were starting work on our Deal I cosmic ray instrument immediately, that Dean Slaughter and John A. Collins would be working

CHAPTER 8 • GO! JUPITER C, JUNO, AND DEAL I 229

with me, and that they were listening in on the line. Those senior engineers in Henry’s New Circuit Elements Group did much of the work in constructing my instruments, testing them, and integrating them into the satellite. They were to become among my closest and most highly regarded associates during my stay at JPL.

Henry’s and my conversation on that day covered some of the preliminary work that they would accomplish before my arrival. As their first priority, they were procuring a supply of the Geiger-Muller (GM) counters. They had already talked to Anton Laboratories, manufacturer of the counters, and the first shipment had been promised for delivery in about a week. His staff were also starting to build a test model of the 700 volt high-voltage supply to power the Geiger counter. Following our discussion, I sent them regulator tubes, transistors, and a transformer so they could breadboard the high-voltage power supply according to the design worked out at the U. S. Army Signal Corps Engineering Laboratories. I immediately called George Hunrath there to coordinate their supply of additional parts kits. Having just made another design revision to improve the voltage regulation, they were nearly ready to ship the kits for the flight instruments.

My time from then until I left Iowa City in mid-November was fully consumed by continuing project coordination and in preparing for my family’s move.

Security was a major concern. Whereas the Vanguard project had been conducted openly, the program to tool up the Army’s Jupiter C satellite launch vehicle was shrouded in secrecy. Recurring delays in the Vanguard launch schedule, the public outcry following the successful USSR Sputnik launches, and the strong element of risk in undertaking the Jupiter launch on such a short schedule made the Army leadership gun-shy, and they chose to operate outside the public spotlight.

That was strongly reinforced on 6 December 1957 following the spectacular and highly publicized launch failure ofVanguard Test Vehicle 3 (TV-3). On that memorable day, with the whole world watching on live television, we were expecting to witness the launch of the first U. S. satellite. Instead, the first stage rocket lost thrust after two seconds, the entire vehicle collapsed into a huge fireball, and the still-bleating satellite spilled onto the ground. The press had a field day!

With that grand public humiliation, the Army became absolutely insistent that de­tails of the Deal project be kept under wraps until a successful launch was achieved. General Medaris at Huntsville strove to make the launch preparations look like just another routine Redstone missile test. Even in highly classified cables between Huntsville and JPL, the launch vehicle was referred to simply as Missile 29. Key personnel whose movements might give away the secret moved under elaborate decoy plans. At the Cape, erection of the booster and the upper stages was particularly sen­sitive, as that could be observed from the public beaches. Thus, the upper stages were

OPENING SPACE RESEARCH

Подпись:covered with canvas for their predawn move from the assembly area to the launch pad. The working platforms and shelters on the gantry were kept in place around the upper portion of the assembly throughout the prelaunch preparations to screen it from the beach observers.

Medaris warned, “I cannot overemphasize the importance of these decoy plans and the absolute necessity of covering this launching as a normal test of a Redstone missile, and I desire it well understood that the individual who violates these instructions will be handled severely.”21

Parenthetically, I never saw that directive and was not fully aware of the force of those instructions at the time, understanding simply that we were not to talk about the project outside of our close circle of coworkers. I came close to violating security on one occasion. During a trip from Pasadena to the Cape (it was for either the Explorer II or III launch), our plane was diverted from Orlando to Tampa because of weather. JPL arranged for a car to drive us from Tampa to Cocoa Beach. During that long drive, believing that all in the car were JPL employees, I began to talk about the upcoming launch. I was immediately interrupted by one of the occupants and told to be quiet. Apparently, the driver was not an appropriately cleared JPL employee.

Thus, I had to be discreet in preparing for our move to California. All conversations related to the move had to be cloaked in oblique terminology. I was permitted to include only the head of the Physics Department in any detailed discussions. Ernie Ray was the acting department head in Van Allen’s absence, and we worked closely together on the details. I withdrew from my university classes without being able to give my professors or classmates any cogent reason. Likewise, we withdrew Barbara from her kindergarten class. My Dad, by now an avid satellite program enthusiast and publicist, had been interviewing me frequently on his daily radio programs and had otherwise kept his audience fully informed of our progress. Suddenly, I had to cut off his source of information, and he could mention only that my family was making a sudden move to California.

The other media, especially the local newspapers, had been carrying extensive coverage of our instrument development. They, too, were suddenly cut off from further information. I had to cancel all plans for public appearances. Rosalie and I made hasty arrangements to move out of our house for several months without being able to tell our landlady the reason. Fortunately, our neighbors when we lived in married student housing at Finkbine Park, Gene and Charlotte Boley, needed a house until Gene graduated in February. Thus, they stayed in our house for the initial period of our absence.

The Defense Department made a guarded public announcement on Friday, 8 Novem­ber, of the president’s decision to launch a satellite with the Army’s vehicle.

CHAPTER 8 • GO! JUPITER C, JUNO, AND DEAL I 231

The formal enabling teletype message reached ABMA the next day. That official go-ahead started the countdown clock. Von Braun had promised that the satellite would be launched within 90 days of approval. Because of scheduling considerations for Cape Canaveral’s downrange facilities, the launch actually had to be planned for 29 January, only 80 days after the approval.

From then on, it was a race against the clock. Work on the instrument repackaging progressed briskly at JPL. I gambled by dispatching some of our household furnishings to Pasadena via moving van on 12 November, two days before receiving Van Allen’s final approval for the program change.

A successful Explorer III launch!

On Sunday, 23 March, I left again for Cape Canaveral to help with the final prepara­tions for the Deal IIb launch. I arrived at Cocoa Beach late that evening and checked into the Coral Sands Motel on the Cocoa Beach strip.

Even then, planning for a future satellite instrument was occupying a significant portion of my attention. On the previous Thursday, during a phone discussion with Van Allen, I indicated that (1) after my trip to the Cape for the launch, I would stop at ABMA for further discussions of the new IGY Heavy Payload (eventually orbited as Explorer 7, as discussed in Chapter 14), (2) I had given the ABMA engineers the power and size requirements for our new instrument, (3) I would be sending him a list of parts to be ordered for its construction, (4) the transmitter power output would

CHAPTER 10 • DEAL II AND EXPLORERS II AND III 281

be about 0.5 watt, (5) we discussed the placement of the GM counter in the payload, and (6) we agreed on a scaling factor of 512 to reduce the counting rate of the GM counter to a value that could be conveyed over the telemetry system.

On Monday, after making my way to the Cape installation, I learned that the JPL engineers had experienced continuing payload problems. Several of the instrument packages interrogated themselves spontaneously—the result of the increased receiver sensitivity built into the flight packages, combined with the complex radio frequency environment in the area. The engineers installed additional filters in the payloads to correct that problem. The data tape recorder in Payload III skipped some of its pulses, and the engineers substituted the recorder from Payload IV. Then a wire from the release solenoid in that recorder broke off, and they replaced the solenoid with the one from the unit they had just removed.

Fortunately, the recorder in Payload II worked perfectly all through the prelaunch testing, and that payload was eventually selected for launching. It continued to operate perfectly on launch day, both for recording and interrogation. In orbit, its operation was flawless throughout the satellite’s entire operating lifetime.

It has always disturbed me that getting the mechanical recording system to operate dependably was such a problem. Today, it would be relatively easy to fabricate a much less troublesome nonmechanical system with solid-state components. But in 1958, we were pushing the state of the art. I believe that the system in Explorer III was as good as I could have made it at the time. I have always realized that a considerable amount of luck was involved in getting the system in Explorer III to operate so well in space.

The launch countdown was a reprise of the Deal IIa launch, except that it progressed with even fewer hitches. The preflight check on the satellite began at 5:12 AM EST on Wednesday, 26 March 1958, with the countdown clock at X – 380 minutes. A low-power system cosmic ray background rate check was begun at 8:09 AM. We made an interrogation check of the onboard recorder at 8:47 AM and a frequency check two minutes later. By that time, the launch countdown had progressed to X – 165 minutes. We made another spot check of the satellite payload at X – 50 minutes. At that time, a hold had been prescheduled to provide time to catch up in case there had been delaying problems. There had not been, so we simply relaxed and waited during that hold.

Van Allen was in the Pentagon again for the launch. While I was listening to the progress of the countdown in the blockhouse, Van was following the same voice comments via a communication circuit between the Cape and Washington. He jotted down the essential content of those comments in his field notebook, which has been

OPENING SPACE RESEARCH

Подпись:preserved as part of the collection of Van Allen Papers in the University of Iowa Libraries, Department of Special Collections. They provide an excellent impression of the progress of a typical rocket-launching countdown during that era. The excite­ment of the occasion is evident in the notes, extracted here in slightly abbreviated and paraphrased form. His notes began soon after 11:00 AM, when the count was resumed14:

26 March 1958. 11:05 A. M. EST. Picked up count at X – 50. Surface weather raining, but not to a deleterious extent.

11:30. As scheduled, the count is holding at X – 30.

11:39. [Predicted] values for 24 March are as follows:

Range at Injection: 775 km.

Initial Perigee height: 352 km (219 mi.)

Initial Apogee altitude: 2043 km (1269 mi.)

Time, Liftoff to Cutoff: 152.7 sec.

To Thrust decay: 155.7 sec.

To Separation: 157.7 sec.

To Second Stage: 396-404 sec.

To Third Stage: 404-412 sec.

To Fourth Stage: 412-420 sec.

Time, Liftoff to Injection: 420 sec.

Nominal firing time 12:30.

11:50. The local weather is improving at the Cape.

12:15. The count is X – 17.

All vehicles have left the pad.

12:20. The pad is clear of all personnel.

12:20. Telemetry checks are being made.

X-12 at 12:20 EST.

Holding at X – 10 as of 12:22. The reason for the hold—checking the beat-beat [tracking system] indications.

The count has been resumed at X – 10 as of 12:28 EST.

Cluster spinup has been started.

Range instrumentation has been checked out and no difficulties have been found.

X – 8 at 12:30 EST. The Cluster has reached 350 rpm.

The Cluster is up to programmed speed and the cluster ignition test signal has been checked. X-5 at 12:33 EST.

X – 4 at 12:34 EST.

All instrument panels have been checked out with no trouble.

X-3.

X – 2.

The range is clear at X – 90 seconds.

X – 1 [minute].

– 50 sec.

– 40 sec.

All preparations complete.

– 20 sec.

– 10

Firing command.

At the Cape, I noted that the work to improve the reliability of the satellite command system and onboard tape recorder had paid off, as there were no problems while the upper-stage tub was spinning. I monitored the signals from the Iowa instrument

CHAPTER 10 • DEAL II AND EXPLORERS II AND III package throughout the countdown and gave my OK on instrument performance after cluster spin-up and finally once again just a few minutes before ignition.

Liftoff occurred at about 12:38 PM EST on Wednesday, 26 March 1958, only eight minutes after the originally scheduled time. The rocket disappeared quickly into low cloud cover, and final injection into orbit occurred at about 12:46. As the payload disappeared from radio range, the Doppler shift looked good, signifying a proper velocity in its path away from the Cape.

I was impressed once again by the sense of overwhelming power as the rocket lifted off and accelerated away from us. I was talking on a phone at the time, giving the liftoff details to the NRL receiving station in Hangar S, where they were also tracking it. As soon as ignition occurred, I could no longer hear my own voice at all, even though I was shouting into the phone.

Van Allen’s field notes continue to provide a vivid account of the ascent and injection:

Ignition normal—main stage.

Liftoff.

Exhaust plume looks good.

Tilt program has started.

Beat-beat indications are good.

Acceleration appears normal.

12:39. Following trajectory very closely.

Program and lateral acceleration are good.

Programming of cluster speed proceeding as planned.

Tilting program completed.

First stage cutoff OK.

First stage] launch was normal. Cluster and instrument compartments are now coasting to apex.

During this coasting period, the variable jet nozzles are aiming the upper stages to the pre-calculated angle for injection.

The on board timer is to ignite the second stage at 396 seconds after liftoff.

12:45. All four stages ignited according to all indications.

12:58. Pickering reported Antigua interrogated successfully.

At the Cape, I, too, learned that the down-range Antigua Minitrack station had made a successful interrogation of the onboard tape recorder soon after the final rocket firing. Only a single command transmission was required, a strong signal containing a two second recorder readout was received, and the tape recorder was reset for its first orbit.

A simplified Minitrack station at Johannesburg, South Africa, acquired the low – power signal as the satellite passed overhead, but it was too early to reveal whether or not the satellite had achieved a durable orbit. As in the case of Explorer I, there was a long delay before we could confirm that the payload was in orbit.

The observers at the Pentagon learned that a receiving station at the Naval Or­dinance Test Station near Inyokern, California, received the new satellite’s signal beginning at about 14:34 EST (19:34 UT). I did not have access to that information at the Cape, but I soon learned there that the low-power transmitter signal was received

OPENING SPACE RESEARCH

Подпись:in quick succession at the Microlock stations at JPL (19:35:05 UT), Earthquake Valley (19:37:33 UT), and Temple City (19:39:55 UT). The Vanguard project’s San Diego Minitrack station also received the signal at about the same time that Earthquake Valley received it.

Such conclusive receipt of the low-power transmitter signal at the completion of the first orbit caused great jubilation. Determining whether the full cosmic ray instrument package was operating properly took more time, however.

Only the Vanguard Minitrack stations had the equipment for commanding the readout of the onboard recorder. The only such station on the West Coast was the San Diego station. They attempted repeatedly near the end of that first orbit to interrogate the recorder but were unable to detect any response. Hearing that, my heart sank, as recovery of the onboard stored data was essential if we were to achieve the full objectives of our experiment.

I remained at the listening post at Cape Canaveral to sweat out the next orbit. For the pass at the end of the second orbit, the Minitrack station at Quito, Ecuador, had the primary data recovery and tracking responsibility. They received the low-power signal and made a series of 10 command transmissions but were unable to detect any response of the high-power system. Our hopes were buoyed a bit, however, by a report that the Lima, Peru, station, marginally within range, might have heard a faint response from the data recorder.

I went from the Cape to Huntsville to discuss the new IGY Heavy Payload. It was not until early afternoon on the 27th, in Huntsville, that I was finally able to get in touch with Jack Mengel in Washington and learned with tremendous relief that some of the subsequent attempts to read the in-flight recorder had produced better results. By that time, receipt of onboard recorder data dumps had been reported from 5 of 12 satellite passes. I was ecstatic!

During the telephone conversation with Van Allen on launch day, in addition to the discussion about the new satellite planning, he asked that I arrange to bring one of the Explorer I spare payloads back to Iowa City for further calibration checks.

Van returned to Iowa City from Washington that evening. During a stop in Chicago, he called Jack Mengel and received the same information that I had received at the Cape, that is, that the Quito interrogation had apparently been received at Lima. On the next day, he received the same news that I was receiving at Huntsville, that is, that 5 of 12 attempts had been successful.

The further history of Explorer III operation is picked up in the next chapter.

Explorer III’s final velocity was higher than the planned value, so that the maximum orbital height (apogee) was quite high. We found, however, that the final stage was pitched up by from four and a half to six degrees from the horizontal when it fired, resulting in a somewhat lower minimum height (perigee) than planned. The initial

CHAPTER 10 • DEAL II AND EXPLORERS II AND III 285

calculations at the NRL Computer Center on Washington’s Pennsylvania Avenue indicated that the perigee height might be as low as 60 miles, which would result in only a two week orbital lifetime before the satellite’s velocity would be slowed by atmospheric drag and it would plunge earthward. By the next day, however, a greater accumulation of tracking data and hard through-the-night work by Joe Siry’s orbital computation team yielded more accurate orbit parameters, including a perigee height of 125 miles and an apogee height of 1735 miles. Based on that new information, the orbital lifetime was projected to be from four to six months, plenty of time for us to conduct our experiment.

Parenthetically, with the higher-than-planned apogee, the satellite turned out to be even more helpful in delineating the anomalous high-intensity radiation than it might have been at the nominal heights.

Explorer III with its attached rocket stage, like Explorer I, spanned 80 inches in total length and six inches in diameter. The total weight placed in orbit was 31.00 pounds, of which 10.83 pounds was the satellite instrument, 7.50 pounds was the shell, and 12.67 pounds was the exhausted final rocket stage.

The initial orbit ranged from 116 miles at perigee to 1739 miles at apogee, with an inclination relative to the Earth’s equator of 33.5 degrees. The initial orbital period was 114.7 minutes. The satellite reentered the Earth’s atmosphere on 28 July 1958.

I was anxious to get to Iowa City following the launch, but, as mentioned above, I had to stop at Huntsville to coordinate our efforts on the new IGY Heavy Payload. Arriving there at midday on Thursday, 27 March, I joined a meeting already in progress. Chaired by Josef Boehm, that meeting dealt with many details of the payload’s mechanical structure and electronic systems. The next day, I met with H. Burke to further discuss electronic circuits, and then again with Boehm on overall project planning. By that time, my notes were frequently referring to the IGY Heavy Payload as the Juno

II project. Further details of that meeting and the follow-up work are described in Chapter 14.15

It was finally possible to leave Huntsville for Iowa City at noon on Friday, but I was frustrated en route by tardy plane departures and missed connections that forced me to remain overnight in Chicago. In a rather foul mood, as I sat waiting at the Chicago airport, I wrote a note that reflected considerable frustration at the situation then facing me, in spite of the exhilaration of the successful launch:

Had a successful visit at Huntsville. But it looks like Juno II will be another crash program. The first two packages (SUI’s part) are to be done by May 1. Looks like I’ll have to rush back to Iowa, or at least to the Iowa payroll. What a life!! I’m getting a little weary of this rat race—after over two years of it. It will be nice to settle down to school for a while, if that ever happens.16

OPENING SPACE RESEARCH

Подпись: 286Since the pressure for me to move ahead quickly with the Juno II circuit design was so intense, my stop in Iowa City was fleeting. Arriving there at mid-morning on Saturday, 29 March, I had several conversations with Van Allen and the others. Although the preliminary indications from the Explorer I data were hugely intriguing, my personal preoccupation during that visit was to brief our team on the results of the Juno II planning meeting.

During spare moments on that trip, I continued writing a paper describing the Explorer I instrument.17 I left Iowa City for my return to Pasadena at about noon on Monday, 31 March.

Endnotes

1 George H. Ludwig, Laboratory Notebook No. 57-6, covering 10 September 1957 to 30 June 1958, p. 33. Undated entry. Also see George H. Ludwig, JPL internal memorandum to Milton Brockman, “To prepare package for spin test,” 25 November 1957.

2 George H. Ludwig, Laboratory Notebook No. 57-6, covering 10 September 1957 to 30 June 1958, p. 51. Entry dated 9 December 1957.

3 Walter K. Victor, Inter-office memorandum to JPL Cognizant Engineers, “Deal II Standard Operating Procedures and Scheduling,” 13 December 1957.

4 George H. Ludwig, letter to James A. Van Allen, 22 January 1958.

5 Ibid.

6 Ludwig, Laboratory Notebook No. 57-6, p. 51. Entry dated 9 December 1957. Also George H. Ludwig, Journal covering 19 September 1956 to 23 August 1960. Entry dated 18 February 1958.

7 Ludwig, Journal covering 19 September 1956 to 23 August 1960. Entry dated 20 February 1958.

8 George H. Ludwig, JPL Travel Report, 20 March 1958.

9 Ludwig, Laboratory Notebook No. 57-6, pp. 96-99. Entries dated 3 and 4 March 1958.

10 George H. Ludwig, Journal covering 19 September 1956 to 23 August 1960. Entry dated 5 March 1958.

11 Richard Witkin, “2D U. S. EXPLORER FIRED, VANISHES; ORBIT IS IN DOUBT,” The New York Times, 6 March 1958.

12 The results of these tests were described in H. R. Buchanan, JPL inter-office memorandum to Walter K. Victor, “Deal II ‘Turn-On’ Receiver Sensitivity Tests,” 14 March 1958.

13 Robert L. Choate, “Calibration Records for the IGY Earth Satellite 1958 Gamma,” JPL Publication No. 126, 27 June 1958.

14 James A. Van Allen, field note pad entry dated 26 March 1958, located in “Papers of James A. Van Allen,” University of Iowa Library Archives, Iowa City, Box 384, folder 2.

15 The meetings during that visit are recorded in George H. Ludwig, Laboratory Notebook No. 58-8, covering 2 to 18 April 1958 and 30 June 1958 to 2 January 1959, pp. 136-137. Entry dated 2 April 1958.

16 George H. Ludwig, Journal covering 19 September 1956 to 23 August 1960. Entry dated 28 March 1958.

17 George H. Ludwig, “Cosmic-Ray Instrumentation in the First U. S. Earth Satellite,” Rev. Sci. Instrum., vol. 30 (AIP, April 1959) pp. 223-229. Reprinted in IGY Satellite Report, no. 13 (Wash., DC: Natl. Acad. Sci., January 1961).

My hurried move back to Iowa City

As mentioned earlier, because of the demands of the new IGY Heavy Payload instru­ment development, and so that I could join in the task of processing and analyzing the Explorer I and III data, Van Allen and I agreed that I should return to the Iowa campus as quickly as possible. Adding to the urgency of my return was the grow­ing possibility that an additional project (beyond the IGY Heavy Payload) might be approved and would also have to be conducted on a crash basis. In fact, that project did quickly materialize, culminating in the launch of Explorer IV and the Explorer V launch attempt, as described in Chapter 13.

Upon reaching Pasadena from my Iowa City stopover, I found that Rosalie had everything under control, and two-week-old George was thriving. The week was completely consumed, on the home front, by preparations for our move back to Iowa City and, at the laboratory, on program planning and detailed design work for the Juno II instrument. Rosalie carried most of the burden of preparing the household for the move, closing all of our bank and utility accounts, terminating our house contract, taking Barbara out of school (again), and packing our personal belongings.

My primary occupation during that week was to design the electrical and mechani­cal configuration of our IGY Heavy Payload instrument and to order its GM counters. I also spent time collaborating with other experimenters and engineers on the new project, including Vernon (Vern) Suomi at the University of Wisconsin, Mr. Hanson, who worked for Gerhardt Groetzinger at the Research Institute for Advanced Studies, and H. Burke at Huntsville.

I also completed the steps necessary to terminate my active employment at JPL. Following their suggestion, I remained an inactive and unpaid member of the JPL staff. That was intended to make it easy for me to return there for postgraduation

Подпись: CHAPTER 12 • DISCOVERY OF THE TRAPPED RADIATION
FIGURE 12.4 The author preparing the spare Explorer I satellite for the move back to Iowa City.

It, along with all University of Iowa laboratory equipment and our personal effects, was loaded into a U-Haul trailer and pulled behind our Mercury for the drive home. The scene is at the rear of our temporary residence on Claremont Street in Pasadena.

employment, if that should be my desire. It gave me a prearranged employment option more than two years before I received my Ph. D. degree. As it developed, I accepted postgraduation employment at the newly forming NASA Goddard Space Flight Center, and the staff arrangement with JPL, for which I was very grateful, was eventually terminated.

My return to Iowa posed an interesting problem. I was going off the JPL payroll, and they had no obligation to pay for my return move. As SUI had had no financial involvement in my original move west in November, and since they were simply resuming my previous employment at Iowa City, they had no legal obligation to help in moving my family and household possessions back. Fortunately, since I was transporting all of our laboratory equipment, parts, supplies, and spare Explorer I and III satellite payloads back to Iowa, Van Allen felt justified in paying for my own direct transportation expenses. He also adjusted my salary for the next few months to help compensate me for flying Rosalie and the three children back.

Rosalie’s return airline flight with Barbara, Sharon, and two-week-old George on Saturday, 5 April, was as uneventful as one might hope under the circumstances, as they were able to take a direct flight from Los Angeles to Cedar Rapids, only 30 miles from Iowa City. My parents picked them up at the airport and delivered them to our Rochester Avenue home.

To keep the expense of the move as low as possible, I rented a U-Haul trailer to transport the laboratory equipment, spare Explorer I and III units (Figure 12.4), and our personal effects. I left Pasadena on Monday, driving our Mercury sedan and the rented trailer via a southern route through Arizona, Texas, the Oklahoma Panhandle, and the new Kansas turnpike to avoid the possibility of lingering harsh winter weather

OPENING SPACE RESEARCH

Подпись:in the high Rockies farther north. I arrived in Iowa City on Friday, 11 April, after a very pleasant solo drive. My journal reported of the trip, “No bad weather, beautiful scenery.”

As I drove into Iowa City, events at the cosmic ray laboratory were unfolding at a feverish pace. The IGY was in full swing. Van Allen and our small cluster of students, faculty, and staff were hard at work on an energetic balloon, rocket, rockoon, and satellite IGY research program.

There was great public and scientific excitement about the burgeoning space pro­gram, especially after the national humiliation of losing the distinction of being first in space to the Soviets. Pressures were mounting for capitalizing on the early U. S. successes as quickly as possible with follow-on space programs.

That Saturday, 12 April, Van Allen, McIlwain, and Ray brought me up to date on the current situation. The satellite instruments on both Explorers I and III had been working flawlessly and were providing a growing tide of data. Explorer I had reached the end of its operating life, and its ground station recordings were being converted on a routine basis to strip chart records and columns of numbers. Explorer III was working well, a reasonable rate of successful interrogations was being achieved, and the first data recordings were reaching us. The account of that Saturday meeting, as recorded in my personal journal several days after the fact, contains the following paragraph:

By now a very startling and interesting result has appeared in the data. We have encountered some extremely high counting rates at the higher altitudes, and at perhaps all latitudes within north and south 33 degrees. Present thinking is that they may be due to electron clouds. Counting rates are probably over 4000 per second. This result appears on both Explorers, and there seems to be no doubt as to its existence.33

We decided at that meeting to change our Juno II heavy payload counter configuration to allow us to study the new phenomenon with greater discrimination.

During late March and early April, Van Allen, with active involvement by Carl McIlwain, continued discussions with Wolfgang Panofsky that had begun at the 11­12 March meeting at JPL. A satellite was being considered that would have detectors arranged to make more quantitative measurements, both of the natural radiation that we were observing and of charged particles that might be injected into trapped trajectories by a high-altitude nuclear burst—what came to be known as Project Argus.

A few days before our 12 April get-together, Van Allen conveyed our growing belief in the existence of the high-intensity radiation regions to Panofsky. That was the first revelation of the new discovery to anyone outside our small group of four.

CHAPTER 12 • DISCOVERY OF THE TRAPPED RADIATION 339

That letter and its background and implications are discussed further in the next chapter.

Van Allen was sufficiently confident in our conclusions by mid-April that he discussed them with several IGY program officials, namely, Richard Porter, Hugh Odishaw, Homer Newell, and William Pickering.34 Those calls were made, most likely, on Monday, 14 April.

The U. S. National Committee had recently established a policy for the release of scientific information derived from U. S. satellites in the IGY program.35 It provided that all satellite-derived data should be conveyed to the U. S. National Committee in advance of any public release. Odishaw reminded Van Allen of that policy and admonished him to make no public announcement of the new discovery until a formal IGY release could be arranged. The two agreed, during that conversation, on a release date of 1 May.

Second-generation spacecraft

During the last two years of the 1950s, the space program advanced rapidly, both in terms of the technology and of the science. The growing experience and confidence of the Soviet and American technicians and scientists, combined with the increasing

CHAPTER 14 • EXTENDING THE TOEHOLD IN SPACE

 

Sputnik III

 

1. MAGNETOMETER

 

10, DEVICE FOR MEASURING

THE INTENSITY OF PRIMARY C05MI RADIATION

 

5. MAGNETIC AND IONIZATION

MANOMETERS

6. ION CATCHERS

7. ELECTROSTATIC FLUXMETER *. MASS SPECTROMETRIC TUBE

9. DEVICE FOR THE REGISTRATION OF HEAVY NUCLEI IN COSMIC RAYS

 

Second-generation spacecraft

2. PHOTO-MULTIPLIERS FOR THE

Подпись: 11. PICK-UPS FOR THE REGIS TRATION OF MICROMETERS REGISTRATION OF THE CORPUSCULAR RADIATION OF THE SUN

3. SOLAR BATTERIES

4. DEVICE FOR THE REGISTRATION OF PHOTONS IN COSMIC RAYS

FIGURE 14.5 Sputnik 3, with identification of its major features. Weighing nearly 3000 pounds and measuring nearly 12 feet long and 6 feet in diameter, it was gigantic in comparison with early U. S. satellites. (Courtesy of the National Aeronautics and Space Administration.)

weight-carrying capability of U. S. launch vehicles, led quite naturally to spacecraft of increasing size, capability, and complexity.

Some of the spacecraft of that period are referred to as second-generation space­craft, distinguished by the inclusion of multiple primary instruments that made in­creasingly discriminating measurements. In many cases, instruments were comple­mentary in nature, carefully chosen to address specific questions.

Sputnik3 Sputnik 3, discussed earlier in Chapter 12, with its immense weight and array of scientific instruments, was the first of the second-generation spacecraft. A full-blown automatic scientific laboratory, the Soviets originally planned that it would be carried on their first satellite launch attempt. Problems with payload development and the resulting launch postponements led to the preparation and earlier flight of the simpler Sputniks 1 and 2.

This spacecraft was indeed remarkable. Launched on 15 May 1958, it carried, in a single carrier, more instruments than had been planned for the entire U. S. Vanguard program. Illustrated in Figure 14.5, the spacecraft was designed to investigate the pressure and composition of the upper layers of the atmosphere, the concentration of positive ions, the magnitudes of the electric charge of the Sputnik and of the Earth’s electrostatic field, the magnitude and direction of the Earth’s magnetic field,

OPENING SPACE RESEARCH

Подпись:the intensity of the Sun’s corpuscular radiation, the composition and variation of primary cosmic radiation, the distribution of the photons and heavy nuclei in cosmic rays, and micrometeors.

Sputnik 3 provided a wealth of new information. Reaching higher latitudes than the earliest U. S. Explorers, it traveled through the lower north cusp of the outer radiation belt. It helped put to rest scientists’ early fears that micrometeorites might be dense enough to seriously impede our ventures into space. Important results related to the geomagnetic field, low-energy ions, and electrons in the far atmosphere and near space and related to cosmic rays were obtained from this mission.10

Explorer 6 Explorer 6 was the second highly successful second-generation space­craft, and the first one by the United States. It was a spheroidal satellite with four solar paddles designed to study a wide range of geophysical and astrophysical phenomena. The arrangement of components within the central cylindrical platform is shown in Figure 14.6.

Whereas the Explorer I, III, and IV and Sputnik 1 and 2 orbits all lay within 1800 miles of the Earth’s surface, and therefore barely edged into the high-intensity radiation belts, the highly eccentric Explorer 6 orbit was another matter. It laced through the entire region of high-intensity radiation from 152 to 26,350 miles and from north 47 degrees to south 47 degrees, making 113 passes through the outer belt during its operating lifetime.

The spacecraft was another product of STL. They had already built and launched Pioneer 0 (Thor-Able 1), Pioneer 1, and Pioneer 2—Explorer 6 was an evolutionary extension of that work. All those early STL missions were initiated by the Air Force’s Ballistic Missile Division, when the three armed services were still vying for major roles in space, i. e., before NASA was formed in October 1958 to head the civilian space program.

Explorer 6 represented a major advance in the development of U. S. spacecraft technology and scientific research. Launched on 7 August 1959 by a Thor-Able-3 ve­hicle from Cape Canaveral, it weighed 141 pounds. One of its major objectives was to develop and test technologies that would be needed for deeper space flight, including journeys of millions of miles into interplanetary space. Long-term electrical power generation and data transmission over great distances were major challenges that guided some of the design considerations. Solar power generation coupled with stor­age batteries provided the electrical power. An onboard receiver facilitated Doppler tracking, fired the injection rocket, changed the rate of data transmission, turned on a simplified television system, and performed other functions. Three data transmit­ters were used—one operated intermittently with a five watt output for tracking and digital data transmission. It was designed so that it would be able to drive a 150 watt amplifier on future deep space missions. Two other transmitters radiated continuously

Second-generation spacecraft

FIGURE 14.6 Top and bottom views of the main instrument shelf in Explorer 6. A quasi­hemispheric dome covered the top of this short cylinder, with a truncated dome on the bottom. Extending outside the cylindrical center section shown here were four paddles containing an array of solar cells to serve as the primary power source—the satellite was often referred to as the pad­dle wheel satellite. The cylindrical structure shown here measured 29 inches in diameter, and the overall satellite height was 26 inches. (Courtesy of the U. S. Air Force.)

OPENING SPACE RESEARCH

Подпись:at 100 milliwatts for analog data transmission. Since similar data were conveyed by the digital and analog systems, the older and more proven analog system was used primarily to monitor the performance of the new “Telebit” digital system that fed the higher-power transmitter.

The ambitious scientific program rivaled that of the Soviet Sputnik 3 program, in spite of Explorer 6’s smaller size and lighter weight, through its use of low-power miniature transistor electronics throughout. With this and the first Pioneer mission, the Air Force (and NASA, once it was formed) provided an opportunity for a new group of experimenters beyond those of us associated with the earlier Vanguard and Juno programs. A team under Robert (Bob) A. Helliwell, L. H. Rorden, and R. F. Mlodnosky at Stanford University provided a Very Low Frequency Receiver and studied the whistler phenomenon and radio propagation through the ionosphere. Carl D. Graves of STL studied electron density above the ionosphere by radio propagation measurements from the UHF and VHF transmitters.

Manring and Dubin at the Air Force Cambridge Research Center continued their earlier work by providing an impact microphone-type micrometeorite detector. Fluxgate and spin-coil magnetometers were developed, and their data were analyzed by an STL team that included Charles P. Sonett, Edward J. Smith, Paul J. Coleman Jr., J. W. Dungey, D. J. Judge, and A. R. Sims. They provided new information on the overall structure of the geomagnetic field and of its temporal variations.

A pair of instruments consisting of an ionization chamber and GM counter was provided by a team at the University of Minnesota headed by John R. Winkler and including Roger L. Arnoldy and Robert A. Hoffman. That group produced a set of rather complete contours of constant counting rate and radiation dosages. Interestingly, the contours displayed a shape quite different from those that we had deduced at Iowa at an earlier time. Their work helped to stimulate a period of energetic research during the next few years to better understand the trapping mechanism, injection and decay processes, and effects of solar variability.

A group at the University of Chicago provided a wide-angle, triple-coincidence, semiproportional particle telescope to investigate the solar modulation of cosmic radiation and the origin and structure of the Van Allen belts. That group was headed by John A. Simpson and included Charles Yun Fan and Peter Meyer at Chicago and Wilmot N. Hess and J. Killeen at the Lawrence Radiation Laboratory in California.

A scintillation counter was prepared by a team at STL consisting of Tom Farley, Al Rosen, and N. L. Sanders to examine the energy spectra of electrons and pro­tons. STL also provided an image-scanning television system. It obtained very low resolution pictures of the Earth that were a precursor to later cloud cover-observing instruments.

CHAPTER 14 • EXTENDING THE TOEHOLD IN SPACE 411

An array of instruments was provided by STL to measure satellite orientation and various engineering parameters. Finally, a group of scientists used the orbit data for studies of lunar and solar perturbations, atmospheric drag, and effects due to ellipticity of the Earth’s equator. Those individuals included Yoshihide Kozai and Charles A. Whitney from the Smithsonian’s Astrophysical Observatory; Kenneth Moe from STL; and A. Bailie, Peter Musen, E. K. L. Upton from the Naval Research Laboratory.

Explorer 7 As mentioned in Chapter 10, serious planning for a second-generation U. S. satellite began as early as March 1958, buoyed by the elation over the successful Explorer I launch. It envisioned retaining the Juno I upper-stage arrangement but substituting the larger Jupiter Rocket for the Redstone first-stage booster, thereby substantially increasing the weight-lifting capability.

Initial planning by the Huntsville and Pasadena engineers and Washington officials proceeded at a rapid pace, and an experiment complement was soon identified. The Huntsville and Pasadena crews initially referred to that satellite as the International Geophysical Year (IGY) Heavy Payload, although the name Payload 8 (PL-8) was sometimes used. It was to include a package for continuation of our original Univer­sity of Iowa cosmic ray research objective, and that objective was quickly upgraded to follow up on the radiation belt discovery. The other experiments included a Solar X-ray and Lyman-Alpha Photometry Experiment under Herbert Friedman’s leadership at the Naval Research Laboratory (NRL), a Radiation and Heat Balance Experiment by the University of Wisconsin group consisting primarily of Verner Suomi and Robert Parent, and a Heavy Cosmic Ray Experiment using an ionization chamber developed at the Glenn L. Martin Company’s Research Institute for Advanced Stud­ies in Baltimore, Maryland, under the leadership of Gerhardt Groetzinger. Several engineering experiments were also included.

Primary support for the IGY Heavy Payload during the pre-NASA era was provided by the U. S. National Academy of Sciences, which renamed the embryonic satellite Payload 16 (PL-16). Presumably, that was because it was the sixteenth U. S. mission (both successful and unsuccessful) that carried the IGY banner. NASA, upon its formation, took over responsibility for the project and renamed it Satellite 1 (S-1), or the first in the series of NASA managed satellites. Its final name after launch became Explorer 7.

The original plan was to launch this satellite (seen in Figure 14.7) in mid-1958. Our initial schedule at Iowa called for delivery of a first flight cosmic ray instrument to Huntsville on 1 May. The initial schedule began to slip as the more urgent work on Explorers IV and V began to dominate the attention of everyone at Huntsville, Pasadena, and Iowa City. Further delays occurred as the Huntsville and Pasadena teams shifted to preparation for the Pioneer 3 and 4 shots.

Second-generation spacecraft
A first attempt to launch S-1 did not occur until 16 July 1959. At liftoff, the power supply for the guidance system failed, and the vehicle was destroyed by the range safety officer 5.5 seconds later. Of course, by that time, the vehicle was barely off the ground, and the destruct command spilled the entire load of fuel and oxygen onto the launch pad. An enormous fire resulted, and those of us in the blockhouse remained sealed there for over an hour as the firefighting crew fought to bring it under control. The blockhouse blast door was ultimately opened, and we emerged to see the wreckage of the vehicle and our payload strewn around the area. I recovered the charred and melted remains of my cosmic ray instrument and a few other bits and pieces, which I retained until turning them over to the Smithsonian’s Air and Space Museum several years ago. Ernst Stuhlinger and I also examined a four foot rattlesnake that had been cooked by the conflagration.

Nearly three months elapsed before a second launch could be attempted. That happened on 13 October 1959, with a picture perfect launch of Explorer 7. The new

CHAPTER 14 • EXTENDING THE TOEHOLD IN SPACE 413

satellite was placed in an orbit that ranged from 356 to 667 miles in height, high enough that the satellite is still orbiting the Earth 50 years later. Its orbital inclination was about 50 degrees, carrying the satellite far enough north and south to provide valuable new information about the Earth’s trapped radiation.

Suomi and Parent’s heat balance instrument worked perfectly. It initiated the era of satellite studies of the Earth’s climate. Using both satellite observations of the Earth’s heat balance and atmospheric cooling rates measured by net flux radiosondes, Suomi was able to establish the important role played by clouds in absorbing radiated solar energy. Those observations established that Earth’s energy budget varies markedly due to the effect of clouds, the surface albedo, and other absorbing constituents. Using these instruments, Suomi and his team discovered that the Earth absorbed more of the Sun’s energy than originally thought and demonstrated that it was possible to measure and quantify seasonal changes in the global heat budget.

By the time of the Explorer 7 launch, Gerhardt Groetzinger, originator of the Heavy Cosmic Ray Experiment, had died. Martin A. Pomerantz, of the Bartol Research Laboratory, took over the experiment and published results in several papers.11

The twin-GM counter cosmic ray instrument was developed by the author, with major assistance by Bill Whelpley. Graduate student John W. Freeman calibrated the counters. The experiment’s purposes were to provide a “comprehensive spatial and temporal monitoring of total cosmic ray intensity, the geomagnetically trapped corpus­cular radiation, and solar protons.” It operated for more than 17 months, broadcasting its data on two frequencies: 108.00 MHz and 19.994 MHz. That second transmitter was set to the low frequency, with a relatively high output power level of 0.6 watt, in order to make it easy for widespread participation in data recovery by radio amateurs and other interested persons.

The particle measurements from our instrument were somewhat anticlimactic. By the time the satellite had finally been launched, Explorer IV, also with a high orbital inclination, had already provided key information on the structure of the lower fringes of the radiation belts. More discriminating instruments for mapping the radiation belts, identifying the causative particles, and learning of their energy spectra had been operated on Sputnik 3 and Explorer 6. Furthermore, the wide-ranging orbit of Explorer 6 and deep space trajectories of Pioneers 3 and 4 had extended the observations much farther into space. Nevertheless, the Explorer 7 counters provided good observations of short – and long-term temporal variations over a relatively long period, from launch on 13 October 1959 to early March 1961.

Brian O’Brian joined our group as an assistant professor in August 1959 and became a major player in the Explorer 7 analysis effort.12

Vanguard III Vanguard III, launched on 18 September 1959, used the seventh and last launcher built under Navy aegis for the IGY. Somewhat heavier than

OPENING SPACE RESEARCH

Подпись:the earlier Vanguards due to an improved final-stage rocket, at a bit over 50 pounds, it carried three primary instruments, a magnetometer by Jim Heppner and his group at GSFC to measure the shape and intensity of the Earth’s mag­netic field, an array of micrometeorite and other environmental sensors by Herman E. LaGow and his group at GSFC, and a pair of ionization chambers by Herb Friedman and his group at NRL to measure the Sun’s X-ray and ultraviolet emissions.

The thousands of magnetic field measurements obtained during its 84 day period of operation provided a charting of the Earth’s magnetic field with an accuracy far greater than hitherto achieved.13 Furthermore, the magnetometer’s measurements of very low frequency signals known as whistlers yielded estimates of electron densities in the high atmosphere.

The impact rate of interplanetary matter was highly variable. No penetrations of the satellite’s shell were detected, and the impact rate was found to be low enough so as to present only a minor hazard to future spacecraft. Even at that, analysis of readings from the micrometeorite detectors put the accumulative influx of cos­mic dust impinging upon the Earth at an impressive figure of about 10,000 tons a day.

The experience gained in the Vanguard program led to a long series of Explorer and Interplanetary Monitoring Platforms at the new GSFC in Greenbelt, Maryland, that continued until the recent past. Those craft provided opportunities for scientists who had cut their teeth on Vanguard to continue their work and for a fresh wave of emerging scientists to join in the grand adventure.

Pioneer 5 Pioneer 5 was a continuation by the Air Force, NASA, and STL of the work begun with Pioneers 0, 1, and 2. Its primary purposes were to further develop the technology needed for deep space operation and to make scientific measurements in space at a distance well removed from the Earth’s influence. The structure, solar paddle arrangement, and weight (about 95 pounds) were all generally similar to those of the earlier missions, and the scientific instruments were furnished, by and large, by the same group of experimenters. The previously anticipated 150 watt amplifier was added to provide the radiated power needed for long-distance interplanetary communication.

Pioneer 5 was launched on 11 March 1960 into an orbit around the Sun lying between the orbits of Venus and Earth. Its apoapsis (greatest distance from the Sun following its final orbital injection) was 0.993 astronomical unit (AU) and its peri – apsis was 0.706 AU. It requires 311.6 Earth days for each complete circuit around the Sun.

Data were received from the craft at 64, 8, and 1 bits per second, depending on distance from the Earth and the size of the receiving station antennae. Most of the

CHAPTER 14 • EXTENDING THE TOEHOLD IN SPACE 415

telemetered data were recovered by the radio telescope at Jodrell Bank Observatory in England and by a tracking station in Hawaii. Useful data were received from the spacecraft for 50 days until 30 April, after which telemetry noise and weak signal strength made useful reception impossible. During the operational period, the high-power transmitter was commanded on about four times each day for 25 minutes duration each time. A new distance record was set for radio reception in interplanetary space when Jodrell Bank received a faint but readable signal from over 22 million miles away.

The mission provided the first measurements of the magnetic field in deep inter­planetary space and of its variations. Most notably, it found that the field did not drop to zero at distances well removed from the Earth but that the Sun’s field remained detectable there. It also measured solar flare particles and cosmic radiation in the interplanetary region.

S-46 It must be noted that throughout the rest of the 1950s, both the Soviets and Americans continued to have a substantial number of disappointing launch failures. One of those was the satellite that we built at Iowa as project S-46. The S-46 mission had a special significance for me, as the project served as the subject for my Ph. D. thesis in electrical engineering.14

This was the second spacecraft that was largely university built, following on the heels of the State University of Iowa’s earlier Explorer IV (and failed Explorer V). The scientific objectives were chosen to examine a manageable subset of the many then – prevailing questions about the high-intensity radiation belt structure and composition. For that purpose, the satellite was designed for an orbit with a high eccentricity and inclination.

Specifically, with that satellite, we hoped to achieve the following:

• monitor the intensity structure of the two principal zones of geomagnetically trapped radiation over an extended period to help establish the origins and gross dynamics of the two zones

• study the correlations with solar activity and with various geophysical phenom­ena such as aurorae and magnetic storms

• study the particle composition and energy spectra of the respective components

• make a first exploratory study of the energy flux of very low energy trapped particles by use of zero-wall-thickness detectors

Built with NASA support, with Les Meredith at GSFC serving as the payload supervisor, our Iowa group did the overall mission and spacecraft design and designed and built the scientific instruments. The major working partners were, again, our friends at the Army Ballistic Missile Agency (later NASA’s Marshall Space Flight Center), who built the spacecraft mechanical structure, battery, solar power system,

OPENING SPACE RESEARCH

Подпись: 416FIGURE 14.8 The S-46 satellite payload. The central cylinder was the familiar six inch in diameter instru­ment container, with the detectors in the portion that protrudes from the top of the cubical solar cell array. The black circles on the top plate of the cylinder are some of the openings for the detectors.

and telemetry system. They also handled all aspects of the launch vehicle preparation and launch.

I continue to marvel at the wonderfully pleasant and productive working envi­ronment that existed between our groups. Individuals there with whom I worked closely on this mission, in addition to Ernst Stuhlinger, were Josef Boehm, H. Burke, Charles Chambers, Gerhardt Heller, Hans Kampmeier, Fred Speer, Sam Stevens, Art Thompson, and Hermann Wagner.

Tracking and telemetry reception was to have been done by a network of NASA stations that would have provided about 90 percent recovery from the highly eccentric orbit. My primary interface there was, again, John Mengel at GSFC.

The instruments included a pair of cadmium sulfide solid state detectors, two GM counters in an electron magnetic spectrometer arrangement, and a third GM counter to establish data continuity with the measurements from Explorers I, III, and IV.

CHAPTER 14 • EXTENDING THE TOEHOLD IN SPACE 417

The detectors were designed, prepared, and calibrated by a group of rising students. John Freeman led the CdS detector effort, assisted by Guido Pizzella, James (Jim) D. Thissell, and Carl McIlwain. Curtis (Curt) D. Laughlin took the electron spectrometer assembly, and Lou Frank led the GM counter calibration effort. Wonderfully effective engineering support was provided by students H. Kay McCune, Bill Whelpley, and Donald (Don) C. Enemark.

The mechanical components of the payload were machined in the department instrument shop by Ed Freund, Robert (Bob) Markee, Robert (Bob) Russell, Edward (Ed) McLachlan, and Michael (Mike) McLaughlin, under the general supervision of its leader, J. George Sentinella. Drafting was provided by Ray Trachta, G. G. Lippisch, A. M. Hubbard, and B. W. Fry. Others helping with the project at Iowa included Gene Colter, John Davies, Chuck Horn, Lucille Lin, Wei Ching Lin, Thomas (Tom) Loftus, Bob Wilson, Keith Wilson, and Andrace Zellweger.

The completed satellite payload is shown in Figure 14.8. The launch attempt was made on 23 March 1960, but the assembly consisting of the second-, third-, and fourth-stage rockets did not function properly. Van Allen made a valiant ef­fort to arrange a second attempt, but an additional launch vehicle was simply not available.

Fortunately, comparable instruments and their derivatives were flown successfully on later spacecraft, most notably on Explorer 10 launched on 25 March 1961, Explorer 12 launched on 15 August 1961, Explorer 18 (Interplanetary Monitoring Platform 1) launched on 27 November 1963, and the Eccentric Orbiting Geophysical Observatory 1 launched on 5 September 1964.

Special acknowledgments

My wife, Rosalie (who now prefers the shorter name “Ros”), was an active partner in the events related in this story. I am indebted to her for that enthusiastic participation and for her forbearance and support during the more than ten-year period of preparing this manuscript.

James A. Van Allen, in addition to providing the leadership for much of the program at Iowa, encouraged and helped me in writing this story. Throughout the process of researching and drafting this manuscript, he provided information and commented on portions of the text. Special thanks are due to him for preparing the book’s foreword.

Leslie (Les) H. Meredith, the first graduate student with whom I worked in the Iowa Cosmic Ray Laboratory, introduced me to the art of balloon instrument design and fabrication. He provided substantial previously unpublished technical and anecdotal information about the rockoon expeditions that has been incorporated into this book. He reviewed the full manuscript and provided substantive comments.

Frank B. McDonald, from my first association with him at Iowa in 1953 through our most recent discussions, has been a strong guide, personal booster, close friend, and a major factor in my professional development. He reviewed segments of the manuscript during its preparation and provided important comments on the full draft.

Ros and I developed especially close personal and professional bonds during our university years with Carl E. McIlwain and his wife, Mary; Laurence (Larry) J. Cahill Jr. and wife, Alice; and Ernest (Ernie) C. Ray and Mary. Ernie passed away before I began writing this book, but Mary Ray assisted in relating Ernie’s role. Carl McIlwain and Larry Cahill reviewed portions of the manuscript during its preparation and provided very valuable assistance by reviewing the full text.

Special thanks are extended to Nancy Johnston and Mary McIlwain, who painstak­ingly proofread the full manuscript.

Others, too numerous to list, encouraged me and provided input during the long process of writing this book. Many of them are mentioned in the text. Grateful thanks are expressed to all of them.

Endnote

1 It can be argued that the Space Age started earlier with, for example, the flight of balloons into the high atmosphere in the early Twentieth Century, or the first launch of a V-2 rocket to a height greater than 100 miles in 1946. In this work, I somewhat arbitrarily mark the beginning of the Space Age with the first durable excursion into the region above the Earth s sensible atmosphere, that is, with the launch of Sputnik 1 on 4 October 1957.

Adding rockets to the program

Initial IGY planning envisioned the use of well-established ground-based instruments and instruments carried aloft by balloons. Although rocket-borne instruments were anticipated, they were not initially seen as a major program component. With time, however, the growing potential of sounding rockets for a wider variety of observations was recognized.

As an early step to organize a rocket component within the United States, the Upper Atmosphere Rocket Research Panel (UARRP) established a Special Committee for the IGY with Homer Newell as chairman. As U. S. planning for the IGY progressed, the U. S. National Academy of Sciences, having the official responsibility for overseeing the U. S. IGY program, established a Technical Panel on Rocketry. The UARRP transferred the Special Committee for the IGY to the Technical Panel on Rocketry, and further planning for the rocket program progressed under that umbrella.

The National Academy notified the CSAGI of the U. S. intent to include a sounding rocket program as part of its contribution to the IGY. That led, at the second meeting of the CSAGI at Rome in September and October 1954, to the formation of a Working Group on Rockets under Homer Newell’s chairmanship. With that step, rocket sound­ings became an integral part of the IGY program, and by the time the IGY opened, a formidable program of rocket observations was in place.

Scientists gather to review IGY progress

Official detailed planning for the IGY had been under way for several years, as related earlier. A series of four meetings of the Council of Scientific Unions’ full CSAGI had been held in Brussels (June-July 1953), Rome (September-October 1954), Brussels (September 1955), and Barcelona (September 1956) to provide overall planning for the endeavor. Specialized regional and discipline meetings were set up to plan specific details. One of those discipline meetings was the first CSAGI Conference on Rockets and Satellites, held in Washington, D. C., on 30 September through 5 October 1957. The conference agenda included general reports from all countries having IGY rocket and satellite activities, establishment and meetings of working groups, and the presentation of technical papers. Four working groups were established during the opening plenary session, and those groups worked throughout the conference to prepare specific resolutions for endorsement by the full conference.

OPENING SPACE RESEARCH

Подпись:The official record of the conference is contained in volume 2B of the Annals of the International Geophysical Year.8

I combined participation in that conference with further testing and coordination of our instrument development at the Naval Research Laboratory, as described in the previous chapter. Van Allen and Larry Cahill were on the State University of Iowa (SUI) equatorial and Antarctic shipboard rockoon-launching expedition, so I was the sole conference attendee representing the university. I presented, on behalf of their authors, all four of the SUI papers, one a report of the status of our satellite instrument development by Van Allen and me, two others exclusively authored by Van Allen, and one by Cahill and Van Allen.9

Herbert Friedman, head of the U. S. delegation for the Working Group on Internal Experiments and Instrumentation Program, asked me to participate in the work of his group in Van Allen’s absence. I also participated, for the same reason, in the activities of the Working Group on Rocketry.

Much has been said about the participation and comments by the Soviets at that conference. The Soviet delegation was headed by Anatoly A. Blagonravov. By that time, he had risen to the rank of lieutenant general of artillery. In June 1946, the USSR had set up an Academy of Artillery Sciences, with Blagonravov as head of its Department for Rocketry and Radar. He soon became that academy’s president. In that position, he first seriously considered the development of an Earth satellite in 1948, based partly upon the stimulus of captured German documents that described Eugen Sanger’s antipodal bomber, a piloted, winged rocket that would reach an altitude of 160 miles and skip on the top of the atmosphere halfway around the world.

By the time of the 1957 conference, Blagonravov was a full-fledged member of the Soviet Academy of Sciences and of the Interdepartmental Commission on Interplan­etary Communications. He, along with Leonid Sedov, would act as “front men” for Soviet science at international gatherings for many years to come. A chain smoker, his demeanor was courteous and mild-mannered. In contrast to the two younger men with him in Washington, he appeared very distinguished and professorial, with his shock of white hair. Although his inner intensity showed from time to time, according to Walter Sullivan’s account, “much of the time he wore a thin smile and carried his Russian cigarette tipped upward at a rakish angle.”10

He made the formal USSR report at the opening session on Monday, 30 September. In that report, he spent most of his time outlining plans for 85 to 95 rocket launchings from three sites: (1) Franz-Josef Land, (2) in the Antarctic near Mirny, and (3) between 50 and 60 degrees east longitude. He did include a short statement that satellites would be launched, and, simply, that the onboard experiments would vary. He also attended the Tuesday session of the Working Group on Satellite Internal Experiments and Instrumentation Program and the Thursday afternoon session of the

CHAPTER 6 • SPUTNIK! 165

Working Group on Satellite Vehicles, Launching, Tracking, and Computation. Again, very little information about the Soviet satellite program was forthcoming.

A. M. Kasatkin, the second-ranking Soviet delegate and another highly placed Soviet scientist, attended the meetings of the Working Group on Rocketry. He also attended all four sessions of the Working Group on Satellite Vehicles, Launching, Tracking, and Computation. He provided substantial specific information about the sounding rocket launch sites. The firings from Franz-Josef Land would be conducted from Cheynea Island; those from Antarctica would be conducted from shipboard near the Soviet station Mirny; and the 50 degree to 60 degree east longitude firings would be fired from the Soviet Union from between 50 degrees and 60 degrees north latitude. One action of that working group was to agree on a form to notify all participating countries whenever rockets were to be launched. Kasatkin agreed that they would fully comply with the use of that form with regard to the launching of meteorological rockets, but only information on the instrumentation and containers would be pro­vided for other geophysical rockets. They made no commitment whatsoever about notifications related to satellite launches.

S. M. Poloskov was the third Soviet delegate. He served as vice chairman of both the Tuesday and Wednesday sessions of the Working Group on Satellite Internal Experiments and Instrumentation Program. During those presentations, Poloskov stated that the first USSR satellite would carry 20 and 40 MHz transmitters, and that the question of the frequency to be used in later satellites was open for further consideration. He also stated, in response to a query, that the USSR satellite orbit would probably be highly elliptical so that the height of passage over a given spot would vary markedly. Beyond those two points, he, too, said nothing of specific Soviet satellite plans.

Although there had been some advance work in preparing for the meetings of the working groups, the memberships of those groups were not established until the conference opening, and their agendas were not set until their first meetings. As a result, the focus and activities of the working groups evolved progressively during the conference. The working group resolutions were formed at their meetings, and those resolutions represented one of the primary products of the conference.

The technical sessions, on the other hand, were organized purely for the exchange of technical information about the various national programs. The list of sessions, their chairmen, and the listing of officially submitted technical papers had been announced ahead of time, and preprint copies of those papers were distributed at the opening of the conference.

There had been considerable disappointment before the conference that the Soviets had not offered many technical papers. Only three USSR papers, two on satellite

OPENING SPACE RESEARCH

Подпись:tracking and one on determining the composition and pressure of air at high altitudes from sounding rockets, had been submitted prior to the conference and included on the official program.

However, when the USSR delegates arrived, they introduced 17 additional papers. Copies of all of those papers were hastily made and distributed, but most of the Western attendees were unable to read the Russian texts. Naturally, the attendees were anxious to hear of the Soviet plans, and changes were made in the technical sessions to accommodate the new material. Although those changes do not appear in the IGY Annals’ general coverage of the conference, some of the new information did show up in the individual reports of working group sessions. My sparse notes from the meeting show, for example, that Soviet papers dealing with substantial details of their rocket programs for measuring the structure of the atmosphere and for meteorology were presented and discussed.

As far as their satellite program was concerned, their new papers focused on Earth satellite orbital dynamics, potential experiments, and other generic topics, rather than on specific Soviet plans.

As mentioned, the decision by the Soviets to use the lower 20 and 40 MHz frequen­cies greatly disturbed the American attendees, as the entire internationally coordinated U. S. radio receiving and tracking system had been designed to operate at a frequency of 108 MHz, as agreed at the Barcelona CSAGI meeting a year earlier. That frequency had been chosen to permit more accurate tracking of satellite motion, since the higher frequency signal would have been subject to less distortion as it passed through the atmosphere.1112

In retrospect, one can see that one highly significant hint about the launch escaped us during that week. Most of the Soviet lectures were delivered in Russian, with simultaneous translation into English. Most of the written papers were also in Russian, and those were not translated until well after the conference. During the discussion following the oral presentation of one of the technical papers, a Soviet delegate made a passing comment about the timing for the first satellite launch. The Russian word was translated at the time as soon, which was taken by the listeners to mean soon on the time scale of the IGY. A more accurate translation of the Russian word would have tipped us off that the Soviet launch was imminent, literally, due at any moment. Having missed that subtlety, we did not anticipate that the first launch would occur only a few days later.

A hurried move to California

Early on Friday, 15 November, Rosalie, our two children Barbara and Sharon, and I headed west from Iowa City. The trunk and backseat of our black-and-white 1956 Mercury sedan were bulging with the prototype instrument package, my laboratory notebooks, a myriad of components and tooling for the flight units, meager clothing for the family, and a few kitchen items. Rosalie and I fitted five-year-old Barbara and four-year-old Sharon into “cockpits” formed among our belongings on the backseat. Rosalie, now more than six months pregnant, made herself as comfortable as possible for the more than 1600 mile trip, and we were off.

Interestingly, that was just two days after one of those major decadal life passages, my thirtieth birthday. To celebrate it, I was setting off on another great adventure, full of grand expectations and supreme confidence.

It was a time before modern interstate highways. Our path took us along old U. S. Highway 6 through Iowa, Nebraska, and Colorado into Utah, down central Utah on Highway 89, and along Highway 91 through southern Nevada and California to Pasadena. Passing through Denver on Colfax Avenue, we approached the Rockies. Climbing to the top of the continental divide, we paused just long enough to admire the windswept snow and to take a picture of a shivering Rosalie in front of the sign marking 11,988 foot high Loveland Pass. In Utah, we made a slight detour to drive through Zion National Park. Without pausing to invest in Las Vegas’ chief industry, we descended from Cajon Pass north of San Bernardino in the early Monday afternoon of 18 November. Slightly ahead of schedule, we decided to take what appeared on our map to be a scenic shortcut from Cajon Junction, along Angeles Crest Highway through the San Gabriel Mountains, to La Canada, just outside the JPL gates. That turned out to be a big mistake—we learned to beware of shortcuts! We wound along that tortuous mountain road for hours, finally arriving in La Canada about sundown. It was too late to check in at JPL, so we had a late dinner and settled down in a motel for the night.

OPENING SPACE RESEARCH

Подпись: 232Our California adventure began with a thud in the middle of the night. Sharon fell out of bed! Not surprisingly, she started crying, but no amount of consoling seemed to quiet her. We finally took her in our bed, but her whimpering continued. On a whim, I started feeling her shoulder, where her pain seemed to be centered, and discovered a knot on her collarbone. Realizing that it was probably broken, we did our best during the night to keep her comfortable. Upon rising, our first task was to locate a doctor. After reading the X-rays, he confirmed the broken collarbone diagnosis, took the simple step of binding her shoulder, and we went on our way.

Our next task was to locate a place to stay until we could find rental housing. We finally located a motel that was willing to rent on a day-to-day basis, but at a weekly rate if we stayed as long as one week. It was located on Pasadena’s Colorado Boulevard, about a mile east of the downtown area, and an easy commute to JPL.

I finally entered the JPL gatehouse on early Tuesday afternoon, 19 November. I met immediately with several of the JPL managers and engineers, including Bill Pickering. In addition to being JPL’s director, Bill had also been one of Van Allen’s colleagues from their days of launching instruments on V-2s at White Sands, had a strong interest in the possibilities of research via satellite, and took a strong personal interest in our cosmic ray instrument preparations. We agreed that the most pressing task for me was to turn over all the information and equipment that I had brought from Iowa.

In spite of the press of work at JPL, finding longer-term housing for the family could not be deferred. With help from the JPL housing office, we were lucky to quickly locate a furnished house at 371 Claremont Avenue in north Pasadena. Our landlady, Mrs. Copeland, lived on the upstairs floor. We settled into the main level, with the furniture and kitchen items provided with the house, the few belongings we had brought with us, and the small shipment that was delivered by the moving company soon thereafter.

With my working days and evenings at the laboratory, Rosalie carried most of the burden of setting up housekeeping in Pasadena. After enrolling Barbara in her new kindergarten, she proceeded to organize the house and take care of the family. As mentioned earlier, she was in her third trimester of pregnancy. In spite of growing discomfort, she accomplished miracles and complained very little.

The house worked out well, with the main problem being its lack of adequate heat. Its sole heat source was a gas-fired convection heater in one wall of the main room. That may have been adequate for a normal Pasadena winter, but the 1957-1958 winter was unusually cold. We shivered in sweaters throughout the entire winter, and Rosalie and I worried constantly about the children (especially newly born George during the latter months) as they crawled around on the perpetually cold floor. Nevertheless, we all survived without significant illness.

CHAPTER 8 • GO! JUPITER C, JUNO, AND DEAL I

A hurried move to California

FIGURE 8.3 Aerial view of JPL as it appeared in January 1959. The main road running nearly bottom to top in the picture was later named Explorer Road. The main entrance gatehouse can be seen near the bottom (west end) of that road. The long white-roofed building on the imme­diate right of Explorer Road (Building 111) was the engineering and administration building and contained Director Bill Pickering’s office. Henry Richter’s office (and my desk) was in Building 122, another white-roofed building above and to the right of Building 111 in this view. (Courtesy of NASA/Jet Propulsion Laboratory, California Institute ofTechnology.)

The location was ideal, being only a 10 minute drive from JPL, an even shorter distance from downtown Pasadena, and within walking distance from Barbara’s school.

The JPL, in late 1957 and early 1958, was a closely packed facility containing a combination of old wooden structures and a few newer, more permanent laboratory buildings. It was located at the foot of the mountains at the north end of the Arroyo Seco, about two and a half miles north of Pasadena’s Rose Bowl. Figure 8.3 shows the laboratory as it existed at about that time.

A desk, phone, and several file drawers in Henry Richter’s office served as my laboratory home away from home during my five month stay. In addition to his duties as supervisor of the JPL New Circuit Elements and Stable Oscillator Research Group, Henry served as my direct supervisor and primary interface at JPL. Throughout my stay, he helped me with outstanding competence and diligence, and we became fast friends. Although I sometimes suspected that one of his unspoken duties was to “keep that Iowa scientist out of the rest of the organization’s hair,” his role served everyone well, as he provided a well-defined conduit for my interaction with everyone at JPL.

OPENING SPACE RESEARCH

Подпись:His attention to my needs certainly helped in avoiding any confusion that might have resulted, had I been trying to make demands directly upon the many parts of the JPL organization.

I immediately set about turning over the components and equipment that I had brought in the trunk of our car. There was, of course, the operating prototype of the complete University of Iowa Vanguard cosmic ray instrument. Not only was that package studied carefully by the JPL engineers, but it also accompanied Pickering and others for showings at several press conferences and technical meetings.

Included among the components turned over to JPL were flight-worthy parts that I had procured and pretested at Iowa, including various transistors, diodes, resistors, capacitors, relays, tuning forks, and high-voltage regulator tubes. The equipment even included drilling templates that we had made for fabricating the electronic circuit boards and molds for encapsulating the completed circuits in expanded polyurethane foam.

By 22 November, I was up to speed. Van Allen had returned to the Iowa campus from New Zealand, and I was finally able to brief him on the full chain of events that had occurred since the Sputnik launch. I breathed a huge sigh of relief when he was able to resume his normal responsibilities for coordinating the University of Iowa activities.

It was not until 26 November, eight days after my arrival, that I took time off to go through the normal processing as a new employee. By that time, the Personnel Office had become so insistent that I do so that I had to steal a few hours from work to attend to those formalities.