Category Dreams, Technology, and Scientific Discovery

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 He provided substantial previously unpublished technical and anecdotal information about the rockoon expeditions that has been incorporated into this book.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.

Operations and Data Handling

W

ith the launching of the satellites, the work was only half done. Tracking the new birds, operating them, collecting their data, and processing and analyzing the data were of equal importance. The first several of those tasks are addressed in this chapter, with the data analysis effort being deferred to the next chapter.

As in the case of Chapter 5, the details in this chapter may be beyond the interest of some readers. They are included for those who may have a historical interest, or who may find them useful in their professional work. The more casual reader may wish to skip to the next chapter.

Explorer I operation

It was late evening, 31 January 1958, on the U. S. East Coast when Explorer I was launched. By then, it was early morning of the next day in Greenwich, England. It has been common for those most comfortable with local time to mark the launch date as 31 January, and for those preferring universal time to ascribe it to 1 February. Throughout this book, the 31 January convention is used.

A network of ground stations had been established for the Vanguard satellites that could provide tracking and data telemetry contact at least once each orbit. Starting with that basic array, the network evolved over time to make the coverage increasingly robust.

As an interesting side note, the Soviets did not establish a comparable worldwide network for their early Sputniks, but depended on a combination of stations within the USSR, plus coverage by radio amateurs and other volunteers who supplied some

287

OPENING SPACE RESEARCH

Подпись:tracking information. As mentioned previously, an onboard data recorder was included on Sputnik 3, but it failed before launch and was not repaired. An examination of Soviet papers reporting early results from Sputnik 3 reveals that their scientific data coverage was almost entirely limited to the region 20 degrees to 145 degrees east longitude and 42 degrees to 63 degrees north latitude, that is, over the Soviet Union.1

The ground paths of the first several Explorer I orbits are plotted in Figure 11.1, along with the locations of the ground stations. Although based on sparse initial tracking data, the plot from which this figure was derived was accurate enough for early satellite operation, data acquisition, and processing efforts.2 Orbit number 0 (zero) began with the satellite’s orbital injection off the Florida coast and lasted until it passed its first ascending node, that is, its first northbound crossing of the geographical equator. Subsequent orbits were numbered sequentially as the satellite passed each ascending node.

The first public report of data from the Explorer I cosmic ray instrument was issued about four weeks after the launch.3 Although that report provided an excel­lent summary of the situation as it was known then, it was based on very early and incomplete telemetry data. Much of its content dealing with the character of the cosmic ray data was modified later as data analysis progressed, especially af­ter the first data were recovered from the onboard tape recorder in Explorer III on 26 March.

The final assessment of Explorer I performance was pieced together from a large number of sources that were written over a substantial period. The most authoritative was the full tabulation of Explorer I data that was finally published about three years after the launch.4 In summary: [7]

Подпись: Special Publications Opening Space Research: Dreams, Technology, and Scientific Discovery Vol. 62

Operations and Data Handling

FIGURE 11.1 The ground tracks for the first orbits of Explorer I. This version was released less than 12 hours after the launch and was used for initial operations. Throughout the Explorer I lifetime, the orbit was refined repeatedly, as more and more extensive and accurate tracking data were obtained. Locations of the Explorer I and III tracking, data receiving, and commanding stations have been added to the original plot to indicate their relationship to the orbital tracks. (By Wilbur S. Johnston and the author, after a map produced by the Vanguard Computing Center, Naval Research Laboratory.)

 

Подпись: Copyright American Geophysical Union

OPENING SPACE RESEARCH

Подпись:Unexplainably, the high-power transmitter signal reappeared 11 days later, and sparse but partly readable data were obtained during four days from 24 to 27 February. It disappeared for the final time on 27 February.

There was a partial failure of the high-power transmitter’s antenna. That antenna consisted of four flexible stainless steel cables, with the base of each cable anchored at the high-power antenna insulator; a donut-shaped ring was located between the instrument payload and the fourth-stage rocket motor. The four antenna cables (which we usually referred to as whips) were connected to the transmitter through a phasing harness and served as the active elements of that circularly polarized antenna. The flexible cables can be clearly seen in the drawing of Figure 8.4 and the photograph of Figure 8.5.

It had been expected that those flexible whip elements would be held in ax­ially semirigid positions by the centrifugal force resulting from the spinning of the satellite about its long axis. That did not occur. A Jet Propulsion Lab­oratory (JPL) report dated 21 February stated that the final spin rate during rocket ascent had been as designed, at 750 revolutions per minute (rpm). That was ascertained by direct measurements via the launch vehicle telemetry system and was verified by observation of the amplitude modulation of the high-power signal caused by the normal irregularities in its radiation pattern. An abrupt drop in satellite spin rate to 570 rpm occurred about one second after comple­tion of the final fourth-stage burn. That change in spin rate was accompanied by a concurrent increase in the amplitude of the spin modulation of the radi­ated high-power signal. Those facts pointed to the loss of one of the antenna elements.5

There was a departure of the satellite’s motion in free space from that ex­pected. It showed up as a slow periodic variation in the received signal strengths from the high-power transmitter superimposed on the faster variation resulting from the satellite spin. The period of that slower variation was 6.9 seconds, corresponding to a frequency of 8.4 cycles per minute. The slower modula­tion was first seen following the postburning transient mentioned above and grew in amplitude during the first few orbits. As the slower modulation grew in amplitude, the faster modulation decreased, and it disappeared entirely after several orbits. The new slow modulation continued for the rest of the satellite’s lifetime.

That phenomenon resulted from an unanticipated resonance coupling between the spin rate of the satellite and the free-pendulum oscillation rate of the whip antennas. Those two rates turned out to be nearly the same, with the result that the whip antennas were driven to swing violently back and forth. That bending of the whips resulted in a higher-than-expected dissipation of the rotational energy of the satellite. The original spin was around the body’s lowest moment

CHAPTER 11 • OPERATIONS AND DATA HANDLING 291

of inertia—its long axis. The dissipation of kinetic energy, combined with the law of conservation of angular momentum, caused the satellite to precess in a cone of increasing opening angle. Ultimately, the satellite rotated about its transverse axis, the axis of greatest moment of inertia.6

In simple terms, the satellite was tumbling end over end, rather than spinning around its long axis, as planned.

The unexpected modulation of the transmitted signals did make it somewhat harder for the receiving stations to acquire the signals and receive the data and resulted in more data dropouts than would have occurred otherwise. That situation, fortunately, caused only a minor reduction in the usefulness of the scientific and engineering data.

The announcement

Van Allen’s quiet disclosure of our discovery to IGY program officials in Washington in April, combined with the setting of a date for a public release on 1 May, set in motion a feverish effort in our laboratory to prepare the material in a written form that would be easily understandable and thoroughly convincing.

Most of the rest of the month of April 1958 is a blur in my memory! On the first workday after my return to the Iowa campus, i. e., on the same day that Van Allen was calling Washington to reveal our initial Explorer findings, I left for Huntsville on a several days’ session for Juno II planning. In spite of the exciting revelations from the Explorer data, I had to remain heavily involved in designing our instrument for that new mission. Even that effort was soon trumped by the evolving Argus project, as detailed in the next chapter.

The two new projects had the effect of precluding my attendance at the history­making gathering in Washington, where Van Allen made the trapped radiation an­nouncement on 1 May 1958.

In preparing for the public announcement of the new discovery, Van Allen, Ernie Ray, Carl McIlwain, several other students, and I were engaged during the second half of April in working up the Explorer I and III data and preparing the written paper. Ernie Ray took the lead in the data-processing and plotting efforts, while Van, with help from Carl, Ernie, and Joe Kasper, developed the physical interpretation.

The group’s efforts during that period can best be summarized as (1) refining the plot of the GM counter response to high-intensity radiation; (2) further organizing the Explorer I data and plotting them in a suitable form; (3) plotting Explorer III

OPENING SPACE RESEARCH

Подпись: 340playback data from additional passes, to the extent that tapes were received in time; (4) developing and organizing a cogent physical explanation for the newly observed phenomenon; and (5) preparing the written paper.

Step 1: The Geiger-Miiller counter performance As soon as I arrived back in Iowa City with the spare Explorer I payload, Carl exposed it to his X-ray beam. He hardened the beam (that is, eliminated the lower-energy particles) by placing a three-eighth inch thick brass absorber between the X-ray tube and the GM counter. During his runs, he varied the X-ray tube voltage between 50 and 90 kilovolts and the distance between the X-ray tube and GM counter over a substantial range. Additionally, he used various lead shields between the source and counter. Those three techniques resulted in the production of a very wide range in the X-ray flux at the location of the counter.

Carl produced the semilog plot shown in Figure 12.5 to portray the effect of the Explorer III onboard encoding and recording arrangement.36 Since the linear scale on the vertical axis extended only to 350 counts per second, the entire upper portion (up to an observed actual rate of more than 1400 counts per second) is cut off. The color coding described in the figure caption was used in all subsequent plots of the rates from the satellite’s GM counter.

The flat portion of the curve at 128 counts per second reflects the limitation in transmitted data rate resulting from the encoding associated with the onboard tape recorder in Explorer III. That encoding had been designed to accommodate a rate substantially greater than the maximum rate expected from cosmic rays but was, of course, completely inadequate for the unexpected intense radiation.

Interpreting Figure 12.5 in terms of the flight data, the transmitted rate accurately tracked the GM counter’s incident flux rate at values from zero to 128 counts per second. For rates increasing above 128 events per second, the onboard tape recorder encoding circuits limited the reading to that value. That persisted to a counting rate that would have been about 15,000 counts per second in a zero-dead time instrument. As the flux increased still further, the GM counter pulses piled up at the scaler input, as described earlier, so that more and more of the resulting pulses were incapable of triggering the scaler. In that region, the scaler output rate dropped off, as seen by the sharp decline between 15,000 and 30,000 counts per second near the right of the figure. With still higher incident fluxes, none of the pulses was large enough to trigger the scaler, and its output remained constant, appearing to signify a zero counting rate.

Step 2: Organizing the Explorer I data Most of the earliest work on the Explorer I data was with recordings made at the JPL Microlock stations in California (the first data recordings to arrive in Iowa City). But by 12 April, data were beginning to trickle in from Minitrack stations in South America.

We prepared many different exploratory plots in order to arrive at the most un­derstandable way of summarizing and presenting the Explorer I data.37 From those,

CHAPTER 12 • DISCOVERY OF THE TRAPPED RADIATION

The announcement

FIGURE 12.5 The response ofthe GM counter in the Explorer I spare payload upon exposure to high-intensity X-rays in the laboratory. The pencil trace above the flat region shows the rela­tionship between the incident flux and the observed counting rate from the bare GM counter as measured by a fast laboratory instrument. That curve would have peaked at a value of about 1400 ifthey-axis ofthe plot had been extended that far. The truncation at 128 actual counts per second shows the limiting effect ofthe scaler and encoding circuits. The original plot was color coded, with the ascending portion at the left being blue, the flat top green, and red for the zero level on the right. (Courtesy ofthe University Archives, Papers of James A. Van Allen, Department of Special Collections, University of Iowa Libraries.)

several formats were selected for inclusion in our announcement paper. The first was a simple plot showing the counting rate as a function of height for data received from 25 passes over the California stations.38 It convincingly conveyed a picture, even at the high latitude of those stations, of a rising intensity as the altitude increased, from counting rates of about 25 to over 100 counts per second.

As Explorer I data from locations closer to the equator were examined, the picture began to come into sharper focus. One figure included in the discovery paper showed the geographic latitudes and heights for 12 cases. The counting rates were normal in four of those cases (about 30 counts per second), no counts were observed in seven cases, and one case showed a rate in transition.39 In summary, counting rates below about 400 miles height were normal, regardless of latitude, while those above about 1200 miles appeared to be zero.

OPENING SPACE RESEARCH

Подпись:The third plot of Explorer I data included in the paper showed the results of seven high-altitude passes over the Quito, Ecuador, station and one over Antofagasta, Chile. It showed that four high-altitude passes occurring between about 0 degrees and 20 degrees south latitude produced counting rates that appeared to be essentially zero. Four other passes farther from the equator showed rates that were in transition between normal and zero counting rates.40

Steps 3,4, and 5: The Explorer III data and preparing the paper Turning to the Ex­plorer III data, the first task was to complete and refine the preliminary plot of data from the 28 March San Diego Explorer III readout that Van Allen had drawn in his hotel room. The improved full plot, included here as Figure 12.6, shows the counting rate as a function of time for the complete recorder dump.41 Although this particular plot is undated, it was probably produced at around the time of our pivotal meeting on Saturday, 12 April.

An examination of the final packing lists for the tapes produced for us at NRL revealed that copies of readouts of the onboard recorder were copied at NRL onto five reels of magnetic tape.42 The first reel, dated 4 April, contained data from nine onboard recorder readouts that took place on 27 and 28 March. The list for the second reel, dated 7 April, lists 24 data readouts that took place on 28 March-5 April. List three, dated 9 April, records 24 additional readouts. Packing list four, dated 14 April, lists 27 readouts. The fifth and final list, dated 17 April, lists 17 readouts. It can reasonably be assumed that the tape reels arrived in Iowa City shortly after the dates on those lists. The data on the first reel were of largely unsuccessful readout attempts on the first two days of the satellite life and were not useful. But reels two and three contained a gold mine of information.

Many exploratory data plots were tried. Some of them turned out to be of little use, but two turned out to be invaluable. One was a set of plots of radiation intensity as a function of height above the ground and geographic latitude for nine orbits on 28 through 31 March.43

Those nine charts were merged for the next plot, which highlighted the regions of high radiation intensity as a function of position over the globe. The original chart was color coded but was redrawn in black-and-white form for the disclosure paper.44 The information from that figure, overlaid on a world map, appears here as Figure 12.7. The oddly shaped contour marked “300” over South America and the South Atlantic shows the region where the magnetic field at the Earth’s surface is a minimum, and therefore, where the region of trapped radiation dips closest to the Earth. The figure shows the strong correlation between the highest-intensity radiation readings from the flight data (the bold segments) and the field minimum. The displacement of the bold segments somewhat to the east of the region of minimum magnetic field resulted from the fact that the satellite’s altitude was increasing during those segments and therefore climbing deeper into the radiation region.

CHAPTER 12 • DISCOVERY OF THE TRAPPED RADIATION

The announcement

FIGURE 12.6 Final plot of the San Diego station readout of Explorer III on 28 March. Since the onboard recorder playback occurred in reverse order, the right end of this plot represents the be­ginning of the recording interval, while the left end is the end of the recording and the time of the data readout. The horizontal axis scale at the top of the plot shows the actual universal time of recording. (Courtesy of the University Archives, Papers of James A. Van Allen, Department of Special Collections, University of Iowa Libraries.)

This information provided one of the most convincing arguments that the satellite was observing a region of high-intensity radiation.

Our results were first prepared as a Department of Physics research report.45 Al­though dated 1 May 1958, it was actually completed well before 28 April, since it

Подпись: 344
The announcement
Подпись: FIGURE 12.7 Plot of Explorer III onboard-recorded GM counter data for portions of nine orbits during the four day period 28-30 March 1958. The smooth arcs show the paths of the satellite over the Earth's surface during periods when data were recovered. The light solid segments represent regions in which the counting rates were less than 128 counts per second (where the satellite was below the trapped radiation region). The bold segments represent regions where the counting rates were greater than 15,000 counts per second and therefore reflecting the high-intensity radiation. The segments with ovals are regions in which the counting rates were in the transition region between the low and high values. The dashed segments represent regions of uncertain data. The contours marked 300 and 600 are magnetic field values at the Earth's surface during the 1955 era. (Figure by the author. Original data plots courtesy of the University Archives, Papers of James A. Van Allen, Department of Special Collections, University of Iowa Libraries.)

had reached Washington and was being distributed on that date, marked “Release May 1.”

Our report’s opening included a brief, simple summarizing statement:

[We] conclude that [the blanking of the G. M. counter] is not due to equipment malfunction, but is caused by a blanking of the Geiger tube by an intense radiation field. We estimate that if the Geiger tube had had zero dead time, it would on these occasions have been producing at least 35,000 counts/sec.

It continued with a discussion of the instrument and a more complete summary of the preliminary observations. A third major section contained an interpretation of the data that served to justify the claim that the zero counting rate resulted from exposure to very intense radiation. The concluding section, dealing with implications, dwelt on three major points:

The particles are unlikely to have as much as several BeV of energy each. They must initially be associated with plasmas that seriously perturb the magnetic field at an Earth radius or so.

CHAPTER 12 • DISCOVERY OF THE TRAPPED RADIATION

The energy loss in the residual atmosphere above 1000 km may contribute significantly, if not dominantly, to the heating of the high atmosphere.

There are obvious biological implications of the results.

Throughout this chapter, the singular term region of high intensity radiation, Van Allen Belt, or some variation has been used to describe the discovery. This emphasized the fact that, at first, we thought that we were dealing with a more or less homogeneous phenomenon located in a single large region around the Earth. A little later, after examining the data from Explorer IV and Pioneer 3 (as described in the following chapters), it became clear that there were two distinct regions. Soviet data substantiated that finding. After that, the collective term Van Allen Radiation Belts (plural) was widely used.

An early scorecard

It is obvious from the litany of failed space launches already mentioned that the initial foothold in space was somewhat tenuous. During the first year of the Space Age (4 October 1957 to 4 October 1958), a total of 24 attempts were made by the United States and the USSR to launch into Earth orbit and beyond, of which only seven were successful. Extending the period to encompass the first two years, a grand total of 54 space attempts were made, of which 20 were at least reasonably successful. During those two years, 34 failed attempts ended up as piles of flaming debris on the launch pads, down range, or in premature watery graves.

To provide a broad perspective of those first two years, all the launch attempts during that period are listed in Table 14.1. In spite of the huge disappointments in the failures, all the attempts helped in one way or another to develop the technology and infrastructure that led to the incredibly exciting and productive space program that followed.

Подпись: Copyright American Geophysical UnionПодпись: Name Agency Launch (UT) Weight (pounds) Sputnik 1 USSR 4 Oct. 1957 184 Sputnik 2 USSR 3 Nov. 1957 1121 Vanguard TV3 U.S. Navy 6 Dec. 1957 3.0 Explorer 1 U.S. Army 31 Jan. 1958 30.8 Vanguard TV3 BU U.S. Navy 5 Feb. 1958 3.0 Explorer II U.S. Army 5 Mar. 1958 31.0 Vanguard 1 U.S. Navy 17 Mar. 1958 3.2 Explorer III U.S. Army 26 Mar. 1958 31.0 Sputnik USSR 27 Apr. 1958 2756 Vanguard TV5 U.S. Navy 28 Apr. 1958 3.0 Sputnik 3 USSR 15 May 1958 2926 Vanguard SLV1 U.S. Navy 27 May 1958 21.6 Luna USSR 25 Jun. 1958 ? Vanguard SLV2 U.S. Navy 26Jun. 1958 21.6 Explorer IV U.S. Army 26 Jun. 1958 38.4 NOTSNIK U.S. Navy 12 Aug. 1958 2.3 Thor-Able 1 (Pioneer 0) U.S. Air Force 17 Aug. 1958 83.3 NOTSNIK U.S. Navy 22 Aug. 1958 2.3 Explorer V U.S. Army 24 Aug. 1958 38.4 NOTSNIK U.S. Navy 25 Aug. 1958 2.3 NOTSNIK U.S. Navy 26 Aug. 1958 2.3 NOTSNIK U.S. Navy 28 Aug. 1958 2.3 Luna1958A USSR 23 Sep. 1958 ? Vanguard SLV3 U.S. Navy 26 Sep. 1958 21.6 Pioneer 1 NASA 11 Oct. 1958 83.3 Luna1958B USSR 12 Oct. 1958 ? Beacon 1 NASA 23 Oct. 1958 8.8 Подпись: TABLE 14.1Подпись:Подпись: Apogee (miles) 583 1131 1583 Подпись: 2421 1739 Подпись: 1158Подпись: 1375Подпись: 70,750Perigee

(miles) Comments

134 First satellite

132 Carried dog Laika

Lost thrust at two seconds 221 First U. S. satellite

Control system malfunction No fourth-stage ignition 406 First Vanguard success

116 First full-orbit data coverage

Failed to reach Earth orbit No third-stage ignition

135 USSR Primary IGY satellite Improper third-stage trajectory

Moon attempt; failed to reach Earth orbit Premature second-stage cutoff 163 Participated in Argus project

Argus; no verifiable orbit Moon attempt; first-stage malfunction Argus; no verifiable orbit Upper stages aimed incorrectly Argus; no verifiable orbit Argus; no verifiable orbit Argus; no verifiable orbit Moon attempt; failed to reach Earth orbit Insufficient second-stage thrust

Moon attempt; distance record but failed to reach Moon Moon attempt; failed to reach Earth orbit Upper stages separated prior to burnout

Подпись: Special Publications Opening Space Research: Dreams, Technology, and Scientific Discovery Vol. 62

Pioneer 2

NASA

8 Nov. 1958

86.2

No third-stage ignition

Luna1958C

USSR

4 Dec. 1958

?

Moon attempt; failed to reach Earth orbit

Pioneer 3

NASA

6 Dec. 1958

13.0

63,590

Moon attempt; failed to reach Moon

SCORE

ARPA

18 Dec. 1958

8737

922

114

Sent Christmas message

Luna 1

USSR

2 Jan. 1959

796

~122 million

~90 million

Passed within 3727 miles of Moon

Vanguard II

NASA

17 Feb. 1959

21.6

1952

347

Satellite wobble degraded data

Discoverer 1

U. S. ARPA

28 Feb. 1959

1363

601

101

First polar satellite; no reentry capsule

Pioneer 4

U. S. NASA

3 Mar. 1959

13.0

~106 million

~92 million

Passed within 37,301 miles of Moon

Discoverer 2

ARPA

13 Apr. 1959

1638

215

149

Capsule ejected but lost in Arctic

Vanguard SLV5

NASA

13 Apr. 1959

22.7

Second stage damaged on separation

Discoverer 3

ARPA

3 Jun.1959

1660

Failed to reach Earth orbit

Luna

USSR

18 Jun. 1959

?

Moon attempt; failed to orbit Earth

Vanguard SLV6

NASA

22 Jun. 1959

22.7

Second-stage propulsion malfunction

Discoverer 4

ARPA

25 Jun.1959

1660

Insufficient second-stage velocity

Explorer S-1

NASA

16 Jun. 1959

91.3

Off course; destroyed by range safety officer

Vostok

USSR

18 Jul. 1959

?

Early unmanned test of satellite for manned flight; failed to reach Earth orbit

Explorer 6

NASA

7 Aug. 1959

141

26,350

152

Highly eccentric orbit

Discoverer 5

ARPA

13 Aug. 1959

1722

174

88.9

Capsule orbited

Beacon 2

NASA

14 Aug. 1959

9.9

First-and upper-stage malfunction

Discoverer 6

ARPA

19 Aug. 1959

1726

527

132

Capsule ejected; recovery failed

Luna 2

USSR

12 Sep. 1959

853

Flight time 34.0 hours; impacted Moon

Transit 1A

ARPA

17 Sep. 1959

262

Third-stage malfunction

Vanguard III

NASA

18 Sep. 1959

50.7

2190

319

Final Vanguard launch

Atlas-Able 4A (Pioneer)

NASA

24 Sep. 1959

?

Failed

Luna 3

USSR

4 Oct. 1959

614

296,100

25,040

Photos of Moon’s far side

 

Подпись: Copyright American Geophysical Union

Setting the Stage at the University of Iowa

B

alloons led the initial forays into near space for scientific research. Victor F.

Hess, during a balloon flight in Austria on 7 August 1912, conclusively showed the extraterrestrial origin of cosmic rays.1 That event marked the beginning of an extraordinary chapter in the history of science, in which balloon-based research played an important role.2

Cosmic rays are nuclear particles that travel at extremely high speed. They originate in extraterrestrial space, probably mostly in supernovae. They consist of protons (hydrogen nuclei), alpha particles (helium nuclei), and lesser numbers of higher – charged atomic nuclei, as well as some electrons and photons. Most of the cosmic rays approaching the Earth collide with atoms and molecules in the upper atmosphere to produce showers of secondary radiation. Because few of the primordial cosmic rays ever reach the Earth’s surface, it is necessary to study them from as high above sea level as possible. Balloons remain an important vehicle for their study.

The use of rockets for this purpose was seriously discussed as early as 1929 when, at a meeting in the home of John C. Merriam, then president of the Carnegie Institution in Washington, D. C., one of the attendees optimistically asserted that if a rocket went more than 50 miles high, above the ozone layer, it would “settle the nature of cosmic rays.”3 In 1931, Robert Millikan at the California Institute of Technology tried to persuade Robert H. Goddard to use the high-altitude rockets that he was developing in New Mexico for cosmic ray research. However, Goddard, having become by that time apprehensive about collaborative arrangements and, as a result, an inveterate loner in his rocket work, shied away from such joint endeavors.4

Even during the development of the V-2 rocket (Vengeance Weapon Number 2) in Germany during World War II (WWII), serious thought was given by its designers

5

6 OPENING SPACE RESEARCH

to using it for high-altitude research and space travel, but those thoughts had to be set aside because of the high wartime priority given to developing the weapon. In fact, project technical leader Wernher von Braun and two other staff members were imprisoned by the German Gestapo for two weeks in March 1944—charged with diverting their full attention from their wartime duties by planning to use rocketry for space travel.

An interesting vignette in that connection was related by Ernst Stuhlinger, one of von Braun’s close associates. In the fall of 1944, he visited his former mentor, Professor Hans Geiger, where he lay near the end of his life in a Berlin hospital. Geiger asked his former student what he was presently doing. Stuhlinger replied, “We are working on a long-range precision rocket which, we hope, will be able one day to fly to the Moon.” Stuhlinger went on to explain that he was working on the guidance system that would make it possible. Geiger’s interest was piqued, and he asked, “Do you think you could put a cosmic ray counter on board? And transmit the pulse signals to the ground? And really measure the cosmic ray intensity at high altitudes, far above the atmosphere?” Stuhlinger replied, “Absolutely, and we will certainly not send any of our rockets into space without some scientific instruments on board!”5

It was not until peace followed WWII that the first scientific instruments were carried aloft by rockets. The vehicles first used for that purpose were the captured German V-2 rockets that had been brought to the United States after the war, along with a cadre of senior German scientists and engineers led by Wernher von Braun. The primary purpose of the U. S. V-2 work was to help jump-start a nascent American rocket program.

Fortunately, the German team, with the support of their U. S. associates, followed through on the promise that the rockets would serve a useful purpose by carrying meaningful scientific instruments. By the end of 1950, approximately 63 V-2s had been launched in the United States, most with an assortment of research instruments. Strong leadership for the developing U. S. research program that employed the V-2s was provided by (in simple alphabetical order) Homer E. Newell Jr. (Naval Research Laboratory, NRL), William H. Pickering (Jet Propulsion Laboratory, JPL), Milton W. Rosen (NRL), Homer Joe Stewart (JPL), John W. Townsend Jr. (NRL), and James A. Van Allen (Applied Physics Laboratory, APL). Those individuals all went on to figure prominently in developing the follow-on research rockets and the first U. S. satellites.

New vehicles for high-altitude research were soon developed in the United States, most notably, the Women’s Army Corps (WAC) Corporal by the JPL in California; the Aerobee, developed jointly by the APL and the NRL; and the Viking, developed by the NRL. By the end of 1950, approximately 10 of the WAC Corporals, 50 of the Aerobees, and 7 of the Vikings had been launched.

A more complete discussion of those rocket developments is contained in Chapter 7.

CHAPTER 1 • SETTING THE STAGE AT THE UNIVERSITY OF IOWA 7

Artificial Earth satellites

Serious thinking about reaching farther from the Earth’s surface for scientific research began immediately after the cessation of hostilities at the end of World War II. As

CHAPTER 3 • THE INTERNATIONAL GEOPHYSICAL YEAR 71

stated earlier, atmospheric sounding rockets were widely employed throughout the post-1945 era. But the need for much longer-term observations at higher altitudes was widely recognized.

Early U. S. thinking about satellites In May 1945, soon after his surrender to Allied troops in Germany, Wernher von Braun summarized his views on the potential of rocket-launched satellites for his U. S. Army captors.12 Discussions of satellite possi­bilities within the Army stimulated Navy interest, where the primary initial emphasis was on the observation of ship movements at sea. In October of that year, the Navy became the first U. S. agency to take a major formal step to evaluate the prospects. Its Bureau of Aeronautics set up a Committee for Evaluating the Feasibility of Space Rocketry. That group soon recommended that the design of an instrumented Earth satellite be started.

In December 1945, the Guggenheim Aeronautical Laboratory (later renamed the Jet Propulsion Laboratory) at the California Institute of Technology was given a con­tract to investigate the relationship between carrier vehicle performance, the weight of a satellite, and the height of its orbit. The first result of the Guggenheim study was to point out that the system initially envisioned was too expensive to be supported by then-extant Navy budgets. At that point, the Navy tried to enlist help from the other services. A meeting was held on 7 March 1946 between Navy and Army officials, but the U. S. Army Air Corps declined the invitation to participate. What the Air Corps neglected to mention was that they were starting their own investigation, and they had no intention of sharing their efforts with the Navy.

In a November 1945 report, Air Corps general Henry H. (Hap) Arnold expressed his belief that a spaceship “is all but practicable today.” The next month, an air corps scientific advisory group stated that long-range rockets were feasible, and satellites were a “definite possibility.” In early 1946, the Air Corps commissioned a very highly classified independent study, partly to demonstrate that they, in addition to the Army Ordnance group and the Office of Naval Research, possessed competence in this arena and were qualified to assume responsibility for military satellite missions. Project RAND (standing for “Research ANd Development”) was set up by the air corps within the Douglas Aircraft Company at its El Segundo, California, plant to undertake that study.

That group achieved the remarkable feat of producing its first (classified) report, Preliminary Design of an Experimental World Circling Spaceship, by 2 May 1946.13 That 324 page report envisioned a 500 pound satellite to be launched by a booster that would use the technology obtained from V-2 experience. In addition to its technical assessments, the report identified a number of potential military missions for such a spacecraft, including observation, attack assessment, communications, weather recon­naissance, weapons delivery, and the technological development of missile guidance. There was also a strong focus in the report on the gathering of scientific information

OPENING SPACE RESEARCH

Подпись: 72about the Earth and its near-environment. The report made the following prescient observation:

Though the crystal ball is cloudy, two things seem clear:

(1) A satellite vehicle with appropriate instrumentation can be expected to be one of the most potent scientific tools of the Twentieth Century.

(2) The achievement of a satellite craft by the United States would inflame the imagination of mankind, and would probably produce repercussions in the world comparable to the explosion of the atomic bomb.

The problem outlined in the report was that the cost was expected to be $150 million, a prohibitive amount for that era. A year later, RAND presented a new plan for a smaller satellite that it claimed could be launched for $82 million. But a technical evaluation by a Department of Defense group under Clark B. Millikan reviewed both the Air Corps and Navy plans and reported that the identification of some specific military uses would be required before a military development project could be justified. The final nail was driven in the coffin of those earliest efforts to develop an actual military Earth satellite on 15 January 1948, with Vice Chief of Staff Hoyt S. Vandenberg’s delaying statement that “satellites should be developed at the proper time.”14

Thus, by the opening of the 1950s, there was a strong agreement within military circles that satellite launches were possible. Although many of the suggested jus­tifications for launching military satellites did not appear to warrant the high cost, there was an additional factor. Throughout the cold war era, there was an urgent need for intelligence information to assist in assessing USSR military capabilities. U. S. aircraft photoreconnaissance overflights were made, but the Soviets strongly objected to them, the aircraft were detectable by radar, and they were vulnerable to antiaircraft fire from the ground. But, although satellites offered the possibil­ity of reconnaissance from above the range of ground fire and aircraft, U. S. policy makers were seriously concerned that the Soviets would object to them as a sim­ple extension of what they viewed as the warlike lower-level aircraft flights. The Soviets might then be expected to develop an antisatellite capability, and Earth orbit would become just another battlefront, rather than an outpost for a broad range of uses, including peaceful scientific exploration of the universe and the application of space technology to Earth resource management, communications, and other practical uses.

Within the nonmilitary arena, concrete thinking about space flight was also evident immediately following the war. In 1946, staff researchers at the Naval Research Laboratory (NRL) discussed the use of Earth satellites for unclassified high-altitude research. The idea was set aside at that time as being premature—they concluded that the state of the technology was simply not yet available.

CHAPTER 3 • THE INTERNATIONAL GEOPHYSICAL YEAR

The scientist’s hopes for research in space remained only a dream into the early 1950s. The popular press began to help in spreading the word. Although many of the concepts discussed openly did not benefit directly from the military studies because of the latter’s high security classification, an expectant culture of serious space en­thusiasts slowly emerged. The American Rocket Society began pressing the case for the scientific and peaceful conquest of space in the opening years of the 1950s. It established a Space Flight Committee in 1952, with Milton W. Rosen of the NRL as chairman. The following year, that committee, after being provided with at least some of the sensitive military information, issued a classified report with details about the kinds of actions that would be required to promote space flight. They suggested that the “National Science Foundation study the utility of an unmanned satellite vehicle to science, commerce and industry, and national defense.” They went on to state that “examples of these research uses might be: for a superior astronomical observatory site; for biological and chemical research utilizing non-gravity conditions; for elec­tronic research utilizing a more perfect vacuum of unlimited volume for microwave research in free space; cosmic ray studies; and sophisticated nuclear research; etc.”

A year later (1954), the Space Flight Committee followed its earlier report with an unclassified report to the National Science Foundation titled “On the Utility of an Artificial Unmanned Earth Satellite.” It stated that a satellite would be one of the most important steps toward advancing the cause of space flight, and that it would also increase the country’s scientific knowledge.

Since the IGY was undertaken as a completely open, purely scientific international effort, its planning involved a huge body of scientists, many of whom were unaware of most of the classified military activity. Those who did know about it were constrained from discussing it in the unclassified IGY planning arena.

The possibility of using satellites for scientific research became much more openly discussed following an event that occurred in 1952. Fred Singer began presenting and publishing a series of unclassified papers that espoused the use of small artificial Earth satellites for scientific investigations.15>16>17>18>19>20 As summarized later by Singer:

Partly stimulated by lectures I gave to the British Interplanetary Society in London in 1951,1 developed ideas for an instrumented earth satellite to carry on the kinds of measurements we hadbeendoinginrockets…. It was quitearadical ideaat the time, whichoffendedthosewho pooh-poohed any notion about working in space as well as those who had already set their aim on manned exploration of the solar system. What I brought to the discussion, mainly, was the notion that instrumentation could be miniaturized and that useful research could be done with a satellite weighing only a few kilograms—even if it survived only for days or weeks….

So was born the MOUSE—the Minimum Orbiting Unmanned Satellite of the Earth—with the help of futurist Arthur C. Clarke and rocket engineer Val Cleaver and some alcoholic conviviality at the Players’ Club near Trafalgar Square. For the next few years, I would try to think of all kinds of experiments that could be done by such a satellite: meteorological observations, including worldwide measurements of ozone; ultraviolet measurements of the

OPENING SPACE RESEARCH

Подпись:Sun and other stars; measurements of incoming interplanetary dust as well as the zodiacal light/solar dust corona; magnetic measurements of ionospheric currents; the use of the satellite lifetime to measure the density of the upper atmosphere; primary cosmic rays, and finally, geomagnetically trapped particles. All these ideas were duly worked out and published in some detail.21

Singer’s implied claim to have originated the idea of a small, instrumented satellite was greeted with discomfort by much of the scientific community. As related by Homer E. Newell in his book on the early history of space science:

Members of the Upper Atmosphere Rocket Research Panel were aware of these [early military satellite] studies, but those who were employees of the military did not feel free to press the issue. As has been seen, the panel recommended only a sounding rocket program to the Academy of Sciences [for the IGY]. But geophysicist S. Fred Singer of the Applied Physics Laboratory, who had been conducting cosmic ray and magnetic field research in sounding rockets, felt under no restraints of military security. From some fairly simple calculations, Singer concluded that it should be possible to place a modest (45-kilogram) satellite in orbit around the Earth, and at every opportunity, he urged that the country undertake to do so. Singer’s conclusions were qualitatively correct, but his outspokenness generated some friction for at least two reasons. First, Singer’s manner gave the impression that the idea for such a satellite was original with him, whereas behind the scenes many had already had the idea, and they felt that Singer had to be aware of this. Muzzled by classification restrictions, they could not engage Singer in debate. Second, being unable to speak out, those who had dug into the subject in much greater depth could not point out that Singer’s estimates overshot the mark somewhat, and that his suggested approach was not as workable as others that couldn’t be mentioned.22

Singer continued to press his ideas for the MOUSE. His next step was also chron­icled by Newell:

Singer gained international attention for his proposal when, in August 1953 at the Fourth International Congress on Astronautics in Zurich, he described his idea for a Minimum Orbital Unmanned Satellite Experiment, which he called MOUSE. MOUSE would weigh 45 kilograms, and would be instrumented for scientific research.

The International Scientific Radio Union, at its 11th General Assembly in The Hague, gave support to Singer’s proposal. At the urging of both Singer and Lloyd Berkner, on 2 September 1954 the Radio Union [International Scientific Radio Union, or URSI] adopted a resolution drawing attention to the value of instrumented earth satellites for solar and geophysical observations. Later that month, on 20 September, the International Union of Geodesy and Geophysics [IUGG] at its 10th General Assembly in Rome adopted an even stronger resolution, actually recommending that consideration be given to the use of small scientific satellites for geophysical research. Both the resolution of the Union of Geodesy and Geophysics and the earlier one of the Radio Union were conveyed to CSAGI, which held its third general planning meeting in Rome shortly after the close of the Geodesy and Geophysics Union meeting.23

The URSI resolution recognized “the extreme importance of continuous obser­vations from above the E-region of extra-terrestrial radiations, especially during the forthcoming IGY.” It went on to state, “URSI therefore draws attention to the fact that an extension of present isolated rocket observations by means of instrumented

CHAPTER 3 • THE INTERNATIONAL GEOPHYSICAL YEAR

Earth satellite vehicles would allow the continuous monitoring of solar ultra-violet and X-radiation intensity and its effects on the ionosphere, particularly during solar flares, thereby greatly enhancing our scientific knowledge of the outer atmosphere.”24

The CSAGI satellite challenge The IUGG quickly followed that resolution with an even stronger formal resolution. It was presented to the CSAGI for action, and on 4 October 1954, the CSAGI passed a very slightly edited version of the IUGG resolution. It read:

In view of the great importance of observations during extended periods of time of extra­terrestrial radiations and geophysical phenomena in the upper atmosphere, and in view of the advanced state of present rocket techniques, the CSAGI recommends that thought be given to the launching of small satellite vehicles, to their scientific instrumentation, and to the new problems associated with satellite experiments, such as power supply, telemetering, and orientation of the vehicle.25

That resolution officially introduced the prospect of artificial Earth satellites into the planning for the IGY program. Exactly three years later, the first satellite was launched.

The U. S. response to the challenge The U. S. response to the CSAGI challenge took some time. Homer Newell’s excellent account of this history reads:

The U. S. National Committee for the IGY gave careful consideration to the proposal during the spring of 1955. Support was not immediately unanimous. Clearly the dimensions of this undertaking would be of a different order from the sounding rockets already a part of the IGY planning. Doubts were expressed over the wisdom of including the project in the IGY. Technical aspects were not the only considerations. There was also the concern about what would be the reaction of people to the launching of an artificial satellite that could easily be viewed as an eye in the sky, could well be accorded some sinister import, perhaps even be equated with some kind of witchcraft. Memories of Orson Welles’s Mars invasion had by no means vanished. Most, however, favored endorsing the project. Joseph Kaplan, chairman of the committee, was especially enthusiastic and jokingly coined the phrase “Long Playing Rocket” for the satellite, by analogy with the long-playing records newly on the market. He suggested that, since sounding rockets had become familiar, the idea of a long-playing rocket would prove less disturbing than the completely new concept of an artificial satellite.26

Eventually, the National Academy of Sciences (as sponsor of the U. S. IGY program) and the National Science Foundation (which provided the money) sought approval of a U. S. Earth satellite program. On 29 July 1955, President Dwight D. Eisenhower announced the decision to launch “small, unmanned, Earth-circling satellites as a part of the U. S. participation in the IGY.” That announcement was made simultaneously in Washington, D. C.; in Brussels, Belgium, at a meeting of the CSAGI in the marble great hall of the Academy Palace; and in the 40 countries participating in the IGY.27 With that announcement, organizing the U. S. program shifted into high gear.

The three U. S. armed forces vied for the assignment to plan and execute the satellite technical program. Through a process described in detail in Chapter 7,

OPENING SPACE RESEARCH

Подпись:that responsibility was ultimately assigned by the Secretary of Defense to the Navy Department on 9 September 1955. In turn, the navy secretary assigned it to the Chief of Naval Research on 27 September, and the director of the NRL was given the task of executing it on 6 October.

The National Academy of Sciences retained the responsibility for policy guidance and for interfacing with the various individuals and organizations of the IGY. That responsibility included the prioritization and selection of the experiments.

In early October 1955, the U. S. National Committee for the IGY established a Technical Panel on the Earth Satellite Program, with Richard W. Porter as chairman. It held its first meeting on 20 October. In late January 1956, Porter asked Van Allen to chair a new Working Group on Internal Instrumentation (WGII). At the same time, he asked William H. Pickering to set up a companion Working Group on External Instrumentation (WGEI). The WGII was concerned with the scientific instruments to be flown, while the WGEI dealt with telemetry and tracking. All three groups undertook their work with great alacrity.

In the final analysis, it is virtually certain that the perceived need to develop a U. S. satellite for military needs served as a significant factor in gaining administration support for the IGY satellite effort. The IGY provided a convenient “open, pure science” cover that helped to ensure that U. S. satellites would be accepted in the international political arena.

Initial official soviet actions Although Soviet interest in space flight was also long­standing, the Soviets were slower to reveal their thinking to the outside world in any formal sense. As mentioned earlier, by the time of the 20 September 1954 meeting of the IUGG, the USSR had not even officially committed to participating in the IGY. By early 1956, that commitment had been made, and their IGY Committee was invited by a special letter from the CSAGI secretary-general, Marcel Nicolet, to consider participating in the rockets and satellites program.

Although the invitation was apparently received with great interest in the Soviet Union, no formal announcement of Soviet plans to launch an Earth satellite was made to the outside world until that fall. On 11 September 1956, Academician Ivan P. Bardin announced to the delegates at the Fourth General Assembly of the CSAGI in Barcelona that the USSR would have a rocket program in the IGY and “would use satellites for pressure, temperature, cosmic ray, micrometeor, and solar radiation measurements.”

It was not until 10 June 1957 that Bardin revealed any further details about the Soviet program to the IGY planners. This was done via letter to the CSAGI Reporter on Rockets and Satellites. In that letter, he mentioned that 125 meteorological rockets would be launched from the Arctic, central USSR, and Antarctic. He also mentioned

CHAPTER 3 • THE INTERNATIONAL GEOPHYSICAL YEAR 77

the satellites again, stating that all of the rocket and satellite launches would study “the structure of the atmosphere, cosmic rays, the ionosphere, micrometeors and meteorites, the physical and chemical properties of the upper astrosphere, and more.” But other details, such as the number of planned satellite launches and their sizes, were not revealed.

Although I did not attend the CSAGI and other early international IGY planning meetings, I did observe and participate in much of the detailed planning for the U. S. IGY satellite program throughout 1956 and 1957. The information above was the limit of my knowledge about the Soviet intentions, and I believe that was true for the majority of civilian scientists involved in the U. S. program. We were largely unaware of the many other indicators of Soviet space activity that are detailed in Chapter 6.

The U. S. Vanguard Satellite Program Once the NRL received the assignment for developing the U. S. satellite in early October 1955, work quickly accelerated. The first substantive outline of the form of the U. S. satellite (by that time known as Vanguard) was presented in late November by Homer Newell to the Technical Panel on the Earth Satellite Program.28 He stated that the NRL concept employed two concentric spheres: an outer one, to be 20 or 30 inches in diameter, and an internal, 12 inch diameter sphere to house most of the scientific instruments. The shape of the outer sphere was chosen to optimize optical tracking and the conduct of scientific experiments related to atmospheric drag. Although it was thought then that the inner container should be spherical to help in controlling the temperature of the internal instruments, later study allowed it to be changed to a cylindrical form that permitted more efficient packaging.29

In that early concept, each of the two spheres was to be pressurized independently with helium, although the satellite was to be able to operate even if pressure in the outer sphere was lost due to punctures by small particles (micrometeorites) that were expected to be present in orbit.

That early design posited a total satellite weight of about 22 pounds, with about 2.2 pounds available for scientific instruments, exclusive of telemetry and batteries. It was stated that the data would be recovered by the Minitrack tracking and telemetry system then being developed at NRL under John T. Mengel’s leadership. The antici­pated periods of usable data reception during passage over each ground station were expected to be from eight seconds to as long as a minute for transits that passed directly overhead.

Some of the U. S. scientists were greatly troubled by the small instrument weight allocation. Van Allen suggested during January 1956 that the project consider making some of the satellites cylindrical in shape, with a length of about 18 inches and a diameter of 6 inches. He referred to this as the Mark II configuration.30 The intent of his recommendation was to make more of the total satellite weight available for

OPENING SPACE RESEARCH

Подпись:scientific instrumentation. It should be noted that the Mark II configuration was similar to the satellites that were ultimately launched as the early Explorers.

Some of Van Allen’s preference for the six inch diameter, no doubt, resulted from our group’s experience with that size for the Deacon-based rockoon instruments. And, although he could not discuss it in the unclassified sessions, it is likely that his preference was also strongly influenced by his knowledge of the highly classified Jupiter C developments then under way at the Army Ballistic Missile Agency at Huntsville, Alabama. Van Allen became aware of those developments as early as 1954, when Ernst Stuhlinger, from the Huntsville team, told him of the possibility that the Army might be able to launch a satellite with its rockets. It had been, in fact, that knowledge that most directly motivated Van Allen to prepare his first (November 1954) proposal for a satellite-borne cosmic ray instrument. That background is described in greater detail in Chapter 7.

Major progress had been made in the Vanguard planning by early February 1956. The diameter of the outer shell was set at about 20 inches, in order that it would fall within the envelope of the third stage of the proposed Vanguard launch vehicle. In addition, the configuration for the inner instrument package was changed to a 3.5 inch diameter cylinder with a variable length, depending on the instruments. The satellite maximum weight had been set at 21.5 pounds, with 2 pounds allocated for the experimenters’ instruments (again, exclusive of telemetry and batteries).31

The issue of the satellite configuration remained open for some time. As late as 30 May 1956, two weight breakdowns were still being carried: one for the spherical form favored by NRL and the other for the cylindrical form preferred by Van Allen and several others.32

The issue was finally settled by a compromise of sorts. The outer shell for all the satellites would remain spherical with a diameter of 20 inches, but the specifications for the inner package were amended to permit either a 3.5 inch or a 6 inch diameter instrument cylinder. Although that did not make as much weight available for the scientific instruments as the Mark II configuration would have provided, the 6 inch instrument configuration did permit more efficient packaging.

The Vanguard satellite hardware Upon receiving approval of the Vanguard project in early October 1955, NRL began working diligently to develop the satellite’s shells, thermal control systems, transistor circuitry, telemetry and tracking systems, and other capabilities that would be required. Progress on those fronts (as well as on the launch vehicle and other Vanguard components) is recorded in a series of 36 detailed reports, the first one dated 13 January 1956. In their third report, dated 29 March 1956, they outlined three satellite designs. They were described as follows:

(1) A minimum-weight satellite containing only Minitrack equipment—the size and

shape would be consistent with the equipment and weight requirements. From a weight

CHAPTER 3 • THE INTERNATIONAL GEOPHYSICAL YEAR

standpoint, it would be preferable to attach the sphere solidly to the third-stage shell and omit spin-isolation bearings and separating devices. The temperature and acceleration effects on the structure and equipment are being investigated. The weight of this satellite might be as little as eight pounds; it would not be more than eleven pounds.

(2) A 20-inch spherical satellite weighting 21.5 pounds—it would contain a Minitrack and telemetering transmitter; temperature, pressure, and erosion gauges; and equipment for the measurement of variations in solar Lyman-alpha radiation. It would be mounted on a bearing to reduce the spin rate, and a separating mechanism would cause the satellite to leave the third-stage case at about five feet per second after burnout.

(3) A satellite which would contain the same instrumentation as (2), but might remain attached to the third-stage case and would have an optimum configuration which has not yet been established.

The final choice of the satellite type will be made at a future date.33

The planning, then, was that the first small (six inch) satellites would be built for test vehicle developmental launches, while either the second or third configuration would serve as the full IGY scientific instrument-carrying satellite.

The small satellite was, in fact, placed atop test vehicles (TVs) TV-3 (unsuccessfully attempted on 6 December 1957), TV-3BU (unsuccessfully attempted on 5 February 1958), and TV-4 (successfully launched as Vanguard I on 17 March 1958).34 The third design listed above was dropped fairly early in the program.

The second design evolved into the configuration that was ultimately used for six all-up IGY launch attempts, beginning with the (unsuccessful) launch of TV-5 on 28 April 1958. The one successfully launched on 17 February 1959 atop Satellite Launch Vehicle 4 (SLV-4) was Vanguard II. Vanguard III, launched on 18 September 1959, employed a more powerful third stage, permitting a heavier satellite.

It has been reported from time to time that the small satellites included in the failed launch attempts in December 1957 and February 1958, and successfully launched as Vanguard I in March 1958, were the result of a last-minute crash effort to get a payload into orbit as quickly as possible after the Sputnik 1 launch. In fact, as stated above, the 6.44 inch diameter satellite was always planned as a part of the Vanguard launch vehicle development program.

The number of failed Vanguard vehicle launch attempts may seem excessive at first reading. But it was not so by the standards of rocket development at that time. The Vanguard, during its full development and operational period, made a total of 14 launch attempts, including both early rocket developmental tests and all-up satellite launch attempts, of which there were eight failures.

By comparison, the V-2 rockets assembled in the United States following World War II experienced 20 failures out of 64 attempts, even though thousands of the operational rockets had been launched from Germany by that time, and the U. S. operations were overseen by the German scientists who had helped design them. The development of the Redstone rocket up to its elevation to operational status included 37 test flights, of which 10 were failures. The Thor IRBM, during its developmental

Подпись: 80OPENING SPACE RESEARCH

Artificial Earth satellites

TABLE 3.1 Early U. S. Launching Scorecard

Year

Successes

Failures

Percentage

Success

1957

0

1

0

1958

7

10

41

1959

11

8

58

1960

16

13

55

1961

29

12

71

1962

52

7

88

1963

38

8

83

1964

57

7

89

1965

63

7

90

1966

73

4

95

testing period, made 10 launch attempts, of which 6 were failures. The first version of the Atlas, designated Atlas-A and consisting of only the main sustainer stage with a planned range of 600 miles, experienced five failures out of eight attempts.

Table 3.1 shows the total number of U. S. space launch attempts during the first 10 years of the Space Era, and the percentage of those attempts that succeeded.

Of the eight satellite instruments that I developed at Iowa before my departure in September 1960, only half were successfully launched into orbit. The success-to – failure ratio improved slowly during the next few years. It was only in 1965 that the space launch success rate reached 90 percent.

The Soviets fared no better. The public perception of their success was better, as they hid their failures until much later. In actual fact, during the years 1957 through 1960, they made 19 launch attempts, of which only 9, or 47 percent, were successful.

Experiment selection Long before the announcement in July 1955 of a plan to launch a U. S. satellite, ideas for space-based investigations had been gestating in many minds. The president’s decision and announcement provided a great stimulus for further thinking. The first concrete experiment proposal was Van Allen’s, dated 28 September 1955, to study cosmic rays.35 It was followed a few weeks later by a proposal by Fred Singer at the University of Maryland to measure the erosion of the satellite’s skin by meteoric dust.36 And the competition for real estate on the enthusiastically anticipated satellite was off and running.

The UARRP, mentioned earlier as being formed during the V-2 rocket-launching era, had continued its activity into the IGY planning period. Following the president’s July announcement, that body quickly made concrete plans for examining potential satellite experiments. They ended their meeting on 27 October 1955 at the Ballistics Research Laboratories in Aberdeen, Maryland, with a decision to hold a symposium to discuss ideas for such experiments. The guidelines that they established for that symposium stipulated that attendance would be limited to panel members and their

CHAPTER 3 • THE INTERNATIONAL GEOPHYSICAL YEAR 81

invitees, that only unclassified materials would be considered, and that the subject matter would be highly constrained to be specific, critically considered, and pertinent within the constraints of the present and near-term projection of technologies. The meeting was open for “plans for physical experiments and observations, theoretical and interpretative matters, and techniques and components of a novel nature, but not space medicine, or the legal and political aspects of the satellite program, or essays dealing with vehicle propulsion and guidance.”37

That symposium took place on the campus of the University of Michigan at Ann Arbor on 26-27 January 1956. The forty-third meeting of the UARRP, it was also billed as their Tenth Anniversary Meeting. It provided an opportunity for scientists to present their ideas for space research in an informal, collegial environment. The proposals for both passive and active satellite-based experiments encompassed a wide span of disciplines, including meteor and interplanetary dust characteristics, air pres­sure and density, hydrogen distribution, meteorological measurements, ionospheric structure, temperature, electron density, electromagnetic propagation, auroral radia­tion, magnetic field, Earth heat transfer, solar Lyman-alpha emission, solar stream particles, ultraviolet stellar magnitudes, and, of course, cosmic rays. Thirty-three of those proposals were later published in book form.38

I accompanied Van Allen to that Ann Arbor symposium. It was a watershed experience for me, as it greatly broadened my perspective of scientific research in general, and of the up-and-coming space program in particular. After that meeting (and my concurrent undergraduate graduation), and in advance of official action to select the scientific experiments to be funded by the IGY program, I began substantive work on developing Van Allen’s instrument, as related in detail in Chapter 5.

Immediately after that meeting, at the beginning of February 1956, the newly formed WGII, under Van Allen’s chairmanship, took over the responsibility for all aspects of the instrumentation to be carried on the IGY satellites, including appraising the many proposals being suggested. Initial members of that group were Leroy R. Alldredge (Johns Hopkins Operations Research Office), M. Ference (Ford Motor Company), Herbert Friedman (NRL), William (Bill) W. Kellogg (RAND), Richard Porter (General Electric), Lyman Spitzer (Princeton University), and Van Allen (as chairman).39,40

Thirty serious proposals were initially considered by the WGII. By the time of its first actual meeting on 2 March 1956, the list to be evaluated stood at 11, with 4 more needing additional clarification. They set about energetically to reduce that list to a priority-ordered list that could be flown on the six launch vehicles being procured. The four criteria on which they settled for ranking the proposals were as follows:

(a) Scientific Importance. This aspect was taken to be measured by the extent to which the proposed observations, if successful, would contribute to the clarification and understanding

OPENING SPACE RESEARCH

Подпись: 82of large bodies of phenomena and/or by the extent to which the proposed observations would be likely to lead to the discovery of new phenomena.

(b) Technical Feasibility. This criterion encompassed evidence for previous successful use of the proposed technique in rockets (or otherwise), apparent adaptability of the instru­mentation to the physical conditions and data transmission potentialities of presently planned satellites, nature of data to be expected, and feasibility of interpretation of observations into fundamental data.

(c) Competence. An assessment of competence of persons and agencies making pro­posals was attempted. The principal foundation for such assessment was previous record of achievement in work of the general nature proposed.

(d) Importance of a Satellite Vehicle to Proposed Work. The nature of each proposal was analyzed with respect to the questions: Is a satellite essential or very strongly desirable as a vehicle for the observing equipment proposed? Or could the observations be made nearly as well or better with balloons or conventional rockets as vehicles?41

Van Allen’s Geiger-Muller (GM) counter cosmic ray proposal was accepted on 12 May 1956 by the U. S. National Committee for the 1957-1958 IGY (see the comments in the foreword). It was placed on the short list of potential payloads for early satellite missions, and initial funding was arranged.

At its second full meeting on 1 June 1956, the WGII produced an initial constel­lation of priority-ordered Earth Satellite Proposals (ESPs) out of those that had been submitted. They were as follows:

ESP-8, Satellite Environmental Measurements, H. E. LaGow, Naval Research Laboratory.

ESP-9, Solar Lyman-Alpha Intensity, H. Friedman, Naval Research Laboratory.

ESP-11, Proposal for Cosmic Ray Observations in Earth Satellites, J. A. Van Allen, University of Iowa.

ESP-4, Proposal for the Measurement of Interplanetary Matter from the Earth Satellite,

M. Dubin, Air Force Cambridge Research Center.42

At a meeting in early December 1956, the WGII converted that list into a somewhat modified group in which some of the initial proposals were combined and several others were added. The complete list included studies in meteorology, geomagnetism, ionospheric physics, cosmic rays, meteorites, and astrophysics. They identified that list as a “hard-core program” of onboard experiments, designated them as priority-A experiments, and set the stage for their funding.

The first priority-A package included the instrument proposed by Herbert Friedman of the NRL to monitor the intensity of the solar Lyman-alpha ultraviolet emission line at 1215.7 angstroms. It employed a straightforward ionization chamber covering the range 1100 to 1400 angstroms. That primary experiment was to be accompanied by a group of measurements to determine the effectiveness of the provisions for controlling the temperature within the satellite and to measure the density of the field of micrometeorites and their effect on the outer satellite surface.

Unsuccessful attempts were made to launch that instrument on 28 April 1958 (Vanguard TV-5), 27 May 1958 (Vanguard SLV-1), and 26 June 1958 (Vanguard

CHAPTER 3 • THE INTERNATIONAL GEOPHYSICAL YEAR 83

SLV-2). That set of instruments was ultimately launched on 18 September 1959 as part of a substantially expanded Vanguard III payload.

The second approved package included Van Allen’s cosmic ray instrument, con­sisting of a single GM counter coupled with onboard data storage to provide coverage over the entire geographic area covered by the satellite. The observation of cosmic ray intensity above the atmosphere was expected to reveal the geographical symmetry of the cosmic ray intensity, and the deviations of that symmetry from that of the Earth’s magnetic field. The instrument was also expected to provide a first measurement of fluctuations in the intensity of the primary cosmic rays in order to study their possi­ble sources and the process by which they reached the Earth. It was envisioned that the satellite information would supplement and extend the ground-based cosmic ray observations also being planned for the IGY. A second instrument on that package was a set of sensitive gauges on the outer skin of the satellite for determining the order of magnitude of erosion due to meteoric impacts. That instrument was proposed by Edward Manring and his group at the Air Force Cambridge Research Center in Massachusetts.

The development of the second Vanguard experiment package is the primary subject of Chapter 5.

The third priority-A package consisted of a proton precessional magnetometer to measure the Earth’s magnetic field at high altitudes and over an extended geographical area. The basic instrument, in a 13-inch diameter sphere, was reduced to flight form under James P. Heppner’s leadership at NRL. That third payload also included a 30 inch diameter inflatable sphere proposed by William J. O’Sullivan at the National Advisory Committee for Aeronautics laboratory at Langley Field, Virginia. That sphere was to be separated from the primary satellite and tracked from the ground to provide a sensitive measurement of the density of the Earth’s atmosphere at much greater heights than hitherto possible.

An unsuccessful attempt was made to launch that two-instrument package on 13 April 1959 (SLV-5). The magnetometer flew later as part of the instrument com­plement on the expanded Vanguard III on 18 September 1959. The inflatable sphere was never flown successfully in that form. A somewhat similar sphere, focused on the original objectives, was eventually launched on 16 February 1961 as Explorer 9. Two Echo satellites, launched on 12 August 1960 and 25 January 1964, used technology developed in that program.

The fourth Vanguard launch vehicle was reserved for a meteorological experiment. Two packages were developed, of which one was to be selected for flight. The first instrument was for the observation of cloud cover over a substantial portion of the Earth’s surface. Developed by William G. Stroud, William Nordberg, and their group at the U. S. Army Corps Signal Engineering Laboratories at Fort Monmouth, New Jersey, it employed two photoelectric telescopes to scan the Earth’s surface as the

OPENING SPACE RESEARCH

Подпись:satellite spun, coupled with an onboard tape recorder for data storage over the entire orbit.

An unsuccessful attempt was made to launch that package on 26 September 1958 (Vanguard SLV-3), and it was successfully orbited on 17 February 1959 as Vanguard II. Unfortunately, its scientific value was limited due to an unplanned wobble in the satellite’s spin due to tipoff by the final rocket stage.

The other meteorological experiment was developed by Verner E. Suomi, engineer Robert (Bob) Parent, and their group at the University of Wisconsin. It employed four specially prepared small spheres supported by rods around the outer equator of the satellite. Those sensors were sensitive to radiation at several different wavelengths to provide a measure of the Earth’s radiation balance, i. e., the net effect of radiation arriving from the Sun and of radiation being emitted from the Earth.

An unsuccessful attempt was made to launch the Wisconsin instrument on 22 June 1959 (Vanguard SLV-6). Although that instrument was never successfully launched as a part of the Vanguard program, it was adapted for the Explorer 7 payload that was successfully launched on 13 October 1959.

Six backup experiments were designated by the WGII, in case problems arose with the development of the primary instruments described above. Those packages were never assigned flight space as part of the Vanguard program, although many of their objectives were ultimately achieved by instruments in different forms on later spacecraft.

A memorable cocktail party: The announcement

It was customary at those conferences for receptions to be hosted to provide fur­ther opportunities for social interaction and technical exchange. The U. S. National

CHAPTER 6 • SPUTNIK! 167

Academy of Sciences started this off with a cocktail party on the opening day. This was a grand and glorious affair, with most of the official delegates and other con­ference participants in attendance. The Americans were unusually gregarious, in no small way a result of the growing publicity accompanying the U. S. satellite program and the expectations of a Vanguard launch within the next few months.

During the week, the official delegates and a scattering of newsmen received invitations for a Friday evening cocktail party at the Soviet Embassy. Those invita­tions read, “The USSR Delegation to CSAGO [sic] Rocket & Satellite Conference request the pleasure of the company of [name] at cocktails on Friday, October 4, 1957 at 6:00 o’clock, [at] 1125 16 Street, NW.” Since Van Allen had initially been expected as one of the formally designated U. S. delegates, an invitation had been made out to “Mr. and Mrs. J. A. Van Allen.” As he was busy launching rockoons in the south Pacific and I was attending the conference as the sole University of Iowa representative, the embassy staff had lined out his name and substituted mine. There­fore, purely by an accident of circumstances, I was able to witness that momentous event.

After the technical meetings for that day ended, I freshened up in my hotel room and walked the short distance to the old Soviet Embassy near Scott Circle. I was warmly greeted and escorted to the grand ballroom on the second floor. During the next half hour or so, most of the other guests arrived. As the assembly gathered, the babble of mingled voices swelled, as small, animated groups formed and reformed around the tables of elaborate hors d’oeuvres and the abundantly stocked bars. Most of the discussions centered on the IGY planning and technical information at the conference, and the three Soviet delegates and many of the embassy senior staff mingled freely with the knots of guests.

Just as the party was reaching full swing—as if upon cue from an unseen master of ceremonies—there was an interruption. Not six feet away from where I stood, Lloyd V. Berkner, CSAGI vice president, and reporter for rockets and satellites, climbed onto a chair and clapped his hands loudly to get our attention. The crowd hushed, and he declared13: