LUNAR IMPACT: A History of Project Ranger

Part I. The Original Ranger

Chapter Five - SUPPORTING THE FLIGHT OPERATIONS

All the while, other JPL engineers were conceiving and building an essential part of Ranger's operational system-the part for tracking the spacecraft’s position, retrieving its engineering and scientific information, and recording that information on the ground. Termed here collectively as support of flight operations, this element, like the spacecraft, was intended for NASA's use in subsequent deep space flight projects. It came in time to consist of three distinct components.

First there was the Deep Space Network composed of the radio tracking, telemetry, and command stations at different points around the earth, a control center from which to direct the activities of these stations, and the earth-based communications network that tied the stations and control center together.

Second, there was the mission operations function; this involved, among other things, computing the space trajectory, receiving and analyzing telemetry data for science and data on the condition and performance of the spacecraft, and generating the commands necessary for Ranger to complete its mission successfully. This activity and the Deep Space Network shared a common control center at JPL.

Third, there were launch operations that involved the preflight testing of the launch vehicle and spacecraft at Cape Canaveral, the actual launch itself, and the immediate tracking support downrange from Cape Canaveral that was needed before. Ranger entered its trajectory in space.

THE DEEP SPACE NETWORK

Ranger's designers recognized and accepted the limited opportunities for flight testing and the potential for improving the reliability of a spacecraft by flying a single configuration repetitively. They also recognized that the requirements for planetary space flight would be more severe than those for lunar flight; orienting the design toward the planetary goal was a logical, forward looking choice. This same reasoning had led to the JPL "bus and passenger" concept, where the functions common to the spacecraft in general were separated and maintained from flight to flight, with changes confined to those "passenger" items peculiar to a given mission. The ground-based systems to support flight operations were thought of in the same way. The design and components for each tracking station were to be made to a single, interchangeable standard. New technology would be introduced only after it had been successfully demonstrated at a test site, and after special recording equipment for a given space mission had been installed at the stations for use during the lifetime of that project. As conceived and implemented, the Deep Space Network * would support all of NASA's unmanned deep space flight projects.

* Originally known as the Deep Space Instrumentation Facility, the name: was later changed to the Deep Space Network when more functions were added. The latter term is used throughout the text to avoid unnecessary confusion.

Withal, the spaceborne and earth-based segments of the Ranger system would operate to a single purpose. Scientific information concerning the moon or the interplanetary environment was to be the end product. The scientific experiments would act as the sensing elements, the launch vehicle and spacecraft as the sensor positioners. The Deep Space Network, in turn, would perform the function of position indicator and sensor output recorder. The Network would also permit more accurate measurement of the position and motions of objects in space and allow the efficient transmission to a spacecraft of a command signal to alter its position. At lunar and planetary distances, very sensitive receivers and powerful transmitters that could be pointed directly at the spacecraft would be required on earth, since the weight and power of the spacecraft continued to be restricted by the modest size of the available launch vehicles. In 1958, when ARPA approved the first Pioneer lunar probes, Eberhardt Rechtin and his associates, Walter Victor, Henry Richter, William Sampson, and Robertson Stevens among them, turned their attention to these new demands.

Eberhardt Rechtin was the man ultimately responsible for providing the radio tracking, telemetry, and command of the Ranger spacecraft. Pickering's former student in electrical engineering, he had graduated from Caltech a class behind Burke and Schurmeier in 1946. When he came up from campus to accept a job at the Laboratory four years later, he held a PhD cum laude in electrical engineering. Positive, forceful, energetic, and enthusiastic, Rechtin quickly established a reputation for his ingenious solutions to technical problems. "He could grasp the ramifications of a complex system and foresee benefits and problems far down the road," a colleague remarked, "when many had difficulty seeing the road itself." In 1958 Pickering named Rechtin Chief of the Telecommunications Division. Two years later, as the programmatic aspects of deep space tracking became evident, he was appointed Program Director of the Deep Space Instrumentation Facility * as well (Figure 32). 1

* The Deep Space Instrumentation Facility was composed of the radio tracking, telemetry, and command stations around the world. It did not include the ground-based communications network which linked them to the control center at JPL. All three elements, the stations, communications network, and the control center, eventually made up the Deep Space Network.


 

Fig. 32. JPL Deep Space Instrumentation Facility Director Eberhardt Rechtin


Under circumstances where a spacecraft far from earth rose and set each day like other celestial bodies, two choices for a deep space radio tracking network presented themselves: First, a single station could be constructed in the United States, and the spacecraft could be interrogated during a single period each day when it came into view. Second, a network of three stations, approximately 120 degrees apart in longitude, could be constructed around the globe. While the second choice would involve possible international political complications, it meant that the spacecraft's transmissions could be received and the craft itself kept in view continuously. Before NASA's creation, in the early months of 1958, Rechtin, his colleagues in the Telecommunications Division, and the engineers charged with design of the Juno IV spacecraft, recommended the latter course, a course which ARPA, and subsequently NASA, approved. 2

This decision eased the difficulty in the design of the thermal control system for the spacecraft, and it helped make more reliable the performance of electronic parts aboard the space machine. Continuous operation of the electronics afforded engineers a constant heat distribution pattern with which to work in designing the thermal control system. Turned on and left on, the electronic components were also found to possess a longer lifetime than equipment operated intermittently. Three radio tracking stations, moreover, when fitted out with both receivers and transmitters, added flexibility to the system's ability to detect malfunctions and to trouble-shoot any time a spacecraft encountered difficulty. Designers of the Soviet deep space radio tracking system selected the first, approach, the single tracking station which would view their craft for a single period each day. Operated intermittently, their interplanetary space machine was designed with an active thermal control system (using convection), with most of the electronics hermetically sealed and pressurized inside a tank (see Figure 26). Early failures of the spacecraft on deep space missions caused the Soviets to move their design and operation more closely to the American model, and to build a second tracking station-though both remained confined within the broad borders of the Soviet Union. 3

Members of a JPL team began examining sites for the global deep space stations in 1958-59. They found the first one in March 1958, near Goldstone Dry Lake in Southern California's Mojave Desert, 160 kilometers (100 miles) east of JPL. The choice considerably facilitated acquisition of the land, since this area lay inside the Army's Camp Irwin military reservation. 4 Team members selected the remaining sites in 1959: Island Lagoon, another dry lake bed, near the Woomera Test Range in East-Central Australia, and, for the last station, a shallow valley near Johannesburg, in the Republic of South Africa. The U.S. Department of State worked out arrangements with the respective governments for NASA to use these overseas sites and for the construction and staffing of the installations. 5 Reporting to the national cooperating agencies, but technically and operationally to the American network management, the indigenous staff at the Australian and South African stations would furnish competent and enthusiastic support to this international venture (Figure 33).


 

Fig. 33. Station Locations in the Deep Space Network


Commercial radio stations use as much as 50,000 watts of broadcast power, but Ranger's small transponder would radiate only 3 watts. To acquire and sort out this signal from among the random galactic background noise and man-made radio signals bouncing around in the earth's atmosphere, the antennas at the tracking sites had to be very large. They also had to be steerable-to point at the spacecraft. Finally, they had to be equipped with extremely sensitive electronics to receive the signal from space, and have sufficiently powerful transmitters to send commands to the spacecraft. Two of Rechtin's colleagues, Robertson Stevens and William Merrick, quite literally "lifted" an antenna design from the radio astronomers to meet these demands. The standard dish was 26 meters (85 feet) in diameter, with an equatorial or polar mount based on astronomical requirements. Both the dish and mounting designs suited the needs of deep space tracking, where rates of movement were significantly less than a degree per second, and where a distant spacecraft looked more like a star in the sky than did such rapidly moving objects as missiles or low-altitude earth satellites (Figure 34). 6


 

Fig. 34. Twenty-Six-Meter Radio Antenna at Goldstone


The astronomical antenna, nevertheless, had to be modified for two-way deep space communications. The engineers added a device that would automatically point the antenna at the spacecraft. Once contact was made using a pre-calculated ephemeris, the radio signal from the spacecraft itself would be used to drive the antenna servo pointing mechanism. The parabolic surface of the antenna would collect and focus the spacecraft signal upon a "feed" (signal collector) above the center of the dish, which in turn would send it to a series of sensitive amplifiers that boosted its intensity and processed the signal. To command or interrogate the spacecraft with the transmitter, the process was reversed. 7

In a further modification, the transmitter and receiver on earth were coupled to a diplexer. The diplexer permitted simultaneous transmission and reception via a single antenna, as shown in Figure 35. Earth-to-space transmission, known as "uplink" to its practitioners, was provided by a 10-kilowatt transmitter operating at 890 megahertz. The Ranger transponder received this radio signal, phase-locked with it, stripped off the uplink command data, multiplied and modulated the basic downlink signal with engineering and scientific telemetry,* and retransmitted it from space to earth at 960 megahertz. This mechanization made it possible to determine the velocity and direction of the spacecraft by comparing the frequency transmitted to the spacecraft to the frequency received on the earth to obtain the doppler shift. The trajectory in space could be accurately determined from the observed "two-way" doppler shift and the position of the antenna with respect to the earth and the spacecraft. 8 The large, steerable, parabolic antennas and continuous two-way communications would become the salient features in all NASA space missions and in military space communications for command and control.


 

Fig. 35. The Deep Space Network-Spacecraft Link


* Telemetry, the technique of transmitting instrument recordings over a radio link, was pioneered during World War II As rockets became more powerful and flew to greater altitudes and ranges, engineers and scientists found it increasingly difficult to gather data about the vehicle and any experiments it carried. Most rockets could not be recovered intact after flight testing with the result that much valuable information was lost. With the advent of radio telemetering systems, however, various functions of the rocket's performance, such as fin position, combustion pressure, skin temperature, missile attitude, and critical voltages in electronic gear, could be monitored in flight and the information relayed by radio to ground receiving stations, where it was recorded for analysis.

NASA approved, funded, and monitored Deep Space Network developments after that agency began to function in late 1958. Edmond C. Buckley, Silverstein's Assistant Director for Space Flight Operations, supervised the effort from the Office of Space Flight Programs. Later, when James Webb established the tracking and operations function as a separate program in 1961, Buckley would become Director of the Office of Tracking and Data Acquisition. A graduate of Rensselaer Polytechnic Institute, personable and articulate, Buckley brought to NASA many years of NACA experience in the development of the Wallops Island Launch Range, where he had been responsible for the tracking and instrumentation associated with free-flight research. Buckley and Rechtin respected and liked one another, and got on well together. No major disagreements marred the planning and creation of the Deep Space Network.

The ground communications network connected all of the deep space tracking stations, the tracking radars downrange from Cape Canaveral, and the control center at JPL. Each station was tied to the control center by two teletype circuits and one voice circuit. Over these circuits came the administrative messages by phone and the time-labeled tracking, telemetry, and scientific data in digital form suited to the computers housed at JPL. 9 The Goldstone station, JPL, and Cape Canaveral would be linked on trunk tie lines in the United States. The small tracking antennas downrange from Cape Canaveral, in Puerto Rico, and on Antigua and Ascension Islands, were connected by means of a submarine cable, as was the deep space station in Woomera, Australia. But while one could talk with colleagues in Australia directly via a submarine cable, such a direct link to South Africa did not exist in the early 1960s. Phone calls and teletype messages to the Johannesburg station had to pass by submarine cable to England, then by trunk line to Spain. From North to South Africa, across an entire continent, the calls were patched together with high-frequency radio links and land lines. Both phone calls and teletype information from Johannesburg to JPL returned in the same manner. A good connection over the radio segments in Africa depended on stable ionospheric. conditions, and disruptions during certain times of the day were common.

On July 4, 196 1, NASA declared the Deep Space Network, including the deep space stations, the ground communications system, and the Right control center at JPL, to be operational. 10 As preflight shakedown tests for Ranger 1 began, however, flight controllers encountered an unforeseen problem with the voice circuits. They found the volume on these lines decreasing precipitously as the date of launch drew near, making it extremely difficult for the controllers at JPL to hear their counterparts in other parts of the net. Checks of the circuits failed to reveal clues to account for this mysterious occurrence. Then one astute observer interrupted his conversation and called out "operator." Several voices answered simultaneously. The volume loss had resulted from an increasing number of operators "listening in" as the moment of launch approached. Thereafter, routings were employed that largely eliminated this difficulty, but it did little to improve the poor quality of the high-frequency radio links across Africa. 11 With commercial communication satellites as yet unavailable to handle this international traffic, the radio links would have to be tolerated until 1968, when a submarine cable connecting South America with South Africa was finally completed.

SPACE FLIGHT OPERATIONS

"Space flight operations" would become the art of conducting, on the ground by remote control, a mission being performed by a vehicle in space. It would include determining and correcting the flight path, analyzing telemetry from space, specifying and issuing commands to the spacecraft, and recording and processing the scientific information from the mission. This specialized activity would also be performed in a single flight control center that served both as the command post and as focus of the deep, space stations, the tracking stations downrange from Cape Canaveral, and the ground communications net that tied these elements together around the world.

In May 1960, detailed studies of a system that could receive, evaluate, and process the data returned to earth, and would permit control of the maneuverable Ranger spacecraft in space, began at JPL. The plans were completed in November, with the aggregate effort designated the Space Flight Operations System. The system consisted of the equipment, computer programs, and groups of technical and operational personnel required to support Project Ranger. Since the activity A involved all of the technical divisions in a combined effort, Systems Division Chief Schurmeier was charged with staffing and directing the new function. He would appoint a Flight Test Director to implement the system. The Test Director would report to him on system matters in general, and to Burke as they affected Ranger in particular. Together, they would work out the details of space flight operations. 12

Just as the Ranger spacecraft was broken down into subsystems, space flight operations were likewise divided into functional components, each reporting to the Flight Test Director. There was to be a Deep Space Station Operations Manager, a Central Computing Facility, a Scientific Data Group, a Data Reduction Group, and a Spacecraft Data Analysis Team. The groups would consist of the cognizant engineers and representatives from the technical divisions who had participated in the development and testing of Ranger's subsystems. They would receive the engineering telemetry and determine and analyze the state of Ranger's health. At the end of 1960 NASA and JPL officials decided to house Space Flight Operations in a control center at JPL, along with the computers and the main technical staff. The building containing the computer facility would be modified to incorporate the operations rooms needed for the Ranger flights. Systems Division Chief Schurmeier named Marshall S. Johnson as Flight Test Director. Johnson had been employed in 1957 as one of JPL’s growing collection of computer wizards. The job of Flight Test Director called for securing the cooperation of all of JPL’s technical divisions involved in the conduct of mission operations, a task of no mean proportions. Johnson's knowledge, persuasive talents, and demonstrated willingness and ability to tackle any assignment-or anybody obstructing an assignment marked him as a natural choice. One year later, in October 1961, he was named Chief of a new Space Flight Operations Section as well. 13

The flight control center took shape in 1961 prior to the flight of Ranger 1. A large room was cleared next door to the computers to serve as the operations center. Blackboards and pinboards, used to post current flight information, lined the walls. Desks and phones were installed for the flight operations teams. Perpendicular to the operations center, another room housed the Science, Command, Spacecraft Data Analysis, and Trajectory groups. With time, Johnson and his crew integrated more rooms, added consoles containing closed-circuit television (allowing each controller to view incoming data in the adjacent rooms), and replaced the blackboards with rear projection screens, where information of a more general nature could be displayed for all to see. Johnson somehow managed to overcome the normal bureaucratic inertia and allegiances, forced people to face problems early, worked around the inevitable manmade complications, and improvised the combination of hardware, "software,"* and people that made space flight operations and the flight control center a functioning reality (Figure 36).

* Software means computer programs and operating routines.


 

Fig. 36. JPL Space Flight Control Center, 1961


The new control facility, make shift at best, was recognized as inadequate to deal with several space flight missions operating simultaneously, a condition projected for post-Ranger lunar and planetary flights. In February 1961, JPL Director Pickering formed a committee to evaluate these future requirements. The committee recommended construction of a new building, devoted exclusively to deep space flight operations. In July, NASA Headquarters approved the plans. 14

This building, the Space Flight Operations Facility, would be completed on schedule and placed in operation in early 1964. As Chief of the Space Flight Operations Section, Johnson oversaw the design and construction of the new control center. Seemingly never wearing a coat, dressed in open shirt and red suspenders, he could be found at work any time. His idiosyncrasies made legends at the Laboratory. NASA Space Sciences Director Homer Newell was among those to acknowledge Johnson's efforts at the dedication ceremony: "Planner, cajoler, threatener, union arbitrator, designer, budget parer, trouble shooter," Newell declared, you name it and he had to do it-and he did." 15

Both the original flight control center and the one that replaced it swelled the number of new installations created expressly for NASA's deep space missions. And all of them, including the Environmental Test Laboratory, the Spacecraft Assembly Building, the Deep Space Tracking Stations, and the Space Flight Operations Facility, were to be employed first in Project Ranger. So, too, were the procedures and facilities devised for launching NASA's Atlas-Agena B vehicles at Cape Canaveral in Florida.

LAUNCH OPERATIONS

NASA's global deep space network and space flight operations could control Ranger once the spacecraft had been injected into its space orbit. But the United States Air Force had first to launch Ranger from the Cape Canaveral Missile Test Annex in Florida, a facility controlled from the Missile Test Center at Patrick Air Force Base, situated a few miles away at Banana River. In 1960 Major General Leighton I. Davis commanded both installations. Davis was a spit-and polish West Point graduate and pilot who made it plain that the new space agency, like the Army before it, was welcome at Cape Canaveral as a tenant, subject to the Air Force regulations that governed all operations there. 16

The Air Force group responsible for actually launching the Rangers from Cape Canaveral was the 6555th Aerospace Test Wing. The 6555th oversaw the checkout of the assembled vehicles on the pad, as well as the launch operations proper. This arrangement was made explicit in July 30, 1960, when NASA issued the Agena B Launch Program Management Organization and Procedures. The same document set forth the organizational framework that NASA had selected for its launch operations. Thus, as with launch vehicle procurement, NASA depended on the Air Force both for the facilities and launch support for Project Ranger. And if the organization to procure the vehicles was complex, the one created to launch them was equally confusing. JPL's Lunar Program Director Cummings and Ranger Project Manager Burke would have to "manage" Ranger launchings through the Air Force launch organization, two NASA Headquarters offices, several subsidiary offices, and the Agena B Coordination Board. This problem would become one of the most difficult with which they had to deal in 1960 and 1961. 17

General Ostrander's Office of Launch Vehicle Programs comprised the first headquarters group. In March 1960, Ostrander had created a Launch Operations Directorate with offices at the Marshall Space Flight Center in Huntsville, Alabama, and at Cape Canaveral in Florida. The German-American Kurt H. Debus, a longtime associate of Wernher von Braun, directed the NASA launch organization. Composed of elements of the Army Missile Firing Laboratory, the Debus Launch Operations Directorate acted as the directing field agency for NASA in all matters concerned with NASA launchings at the Atlantic Missile Range..." Quartered at the Cape, Debus would have direct access to General Davis and, through intermediaries, to von Braun in Huntsville. To assist in coordinating activities and to plan and order the range support equipment needed for Atlas-Agena launches, Debus established a subsidiary office at Cape Canaveral called the NASA Test Support Office. 18

Launch operations also involved Silverstein's Office of Space Flight Programs. Silverstein did not choose to let details of launch operations escape his attention, and in March 1960 he established an organization of his own at the range, known as the Office of Flight Missions. This group served as Silverstein's representative at the Cape, reported directly to him, and acted as the central point of contact for the flight project offices, including Project Ranger. In Silverstein's view, it would further serve to "coordinate range support requirements" and provide "administrative and logistic support to all of the NASA elements." 19

By June 1960, JPL hands at Cape Canaveral complained vigorously that this bifurcated arrangement was unworkable. "Effective action [here]" one declared, "will have to await resolution of the administrative maze." 20 The launch program management organization and procedures document issued in July settled the question of authority. Although it did not reduce the number of groups in the multifaceted organization, it at least defined what each one was and how it related to the others. Meantime, Burke directed Luskin's Agena organization at Lockheed to prepare an engineering plan specifying the tracking, telemetry, data processing, data transmission, and computational facilities and equipment required for the support of Ranger. Eventually known as the Program Requirements Document, it contained the needs for the launch hardware and software, hammered out among the participants.

Resolving these and other launch details dragged on throughout 1960, as proposals bounced back and forth between the various operations offices and the technical panels of the Agena B Coordination Board. Agreement, and the necessary subsequent action, were painfully slow. Under the existing procedures, engineers from Debus Launch Operations Directorate could act as "monitors" of NASA's launch operations, but the 6555th Aerospace Test Wing "supervised" the actual work. In the presence of hesitancy and confusion, Air Force representatives bluntly informed NASA: "Just tell the Air Force what you want, and we will put it in orbit for you." 21

Like Hans Hueter and Friedrich Duerr in the Light and Medium Vehicle Office at Huntsville, the Launch Operations Director Kurt Debus chafed in a role where he held responsibility for activities he could not control. As the new year began, NASA launch operations remained in the hands of the Air Force. NASA Headquarters might not wish to press for more direct control over launch operations, since such a move "could only be looked upon by the Air Force as a lack of confidence on the part of NASA," as one official later commented, 22 but the organization formed in response to that need seemed hardly adequate.

From a low point at the end of 1960, matters improved in early 1961. The Agena B Coordination Board was abolished in January, and in February Burke learned that the Air Force had authorized procurement of the mobile antennas needed to track the Atlas-Agena downrange from Cape Canaveral. At the same time, the Air Force made available Launch Complex 12 and Hangar AE at Cape Canaveral for NASA's Project Ranger. Headquarters had also moved to alter the cumbersome organization for launch operations. The new NASA Administrator James Webb, Deputy Administrator Dryden, and Associate Administrator Seamans argued forcefully that the Air Force might properly furnish the launch vehicles as the procuring agency, but that NASA, as the using agency, should be wholly responsible from launch through the completion of a mission. A NASA-Air Force agreement of July 17, 1961, recognized NASA as a joint tenant with the Air Force at Cape Canaveral and called for the eventual replacement of the 6555th Aerospace Test Wing by NASA's own launch groups over a period of time. 23 And shortly thereafter, in early 1962, the entire NASA launch organization was overhauled. Meantime, however, the first Atlas, Agena, and Ranger had been delivered to Cape Canaveral.


Chapter 4  link to the previous page        link to the next page  Chapter 6

Chapter Five - Notes 

The hyphenated numbers in parentheses at the ends of individual citations are catalog numbers of documents on file in the history archives of the JPL library.

1. JPL Announcement No. 39 from William Pickering to Senior Staff, et al., subject: "Appointment of Program Director and Program Deputy Director of the Deep Space Instrumentation Facility (DSIF), " May 6, 1960 (2-242).

2. Eberhardt Rechtin, "Communications to the Moon and Planets," Proceedings of the Radio Club of America, Volume 42, No. 1, April 1966, pp. 1-8. Rechtin, as events turned out, would later (1967-1970) become Director of ARPA.

3. Soviet Space Programs 1966-70 (Science Policy Research Division and Foreign Affairs Division of the Congressional Research Service and the European Law Division of the Law Library, Library of Congress. Washington: Government Printing Office, 197 1 ), p. 152. The second station had been completed by 1970.

4. "DSN Facility Activation Dates," a JPL chart circa April 1967 (2-141).

5. Cf. Edward M. Walters, The Origins of the Australian Cooperation in Space (Comment Edition, HHN-82. National Aeronautics and Space Administration, May 1969) (5-195). Politics and economics continually impinge on the world of science and deep space radio tracking. In 1975 NASA prepared to abandon the Johannesburg site, partially in response to demands from opponents of apartheid.

6. Rechtin " Communications to the Moon and Planets, " p. 3.

7. Eberhardt Rechtin, Space Communications (JPL Technical Release 34-68. Pasadena, California: Jet Propulsion Laboratory, California Institute of Technology, May 1, 1960).

8. William D. Merrick, Eberhardt Rechtin, Robertson Stevens, and Walter K. Victor, Deep Space Communications (JPL Technical Release 34-10. Pasadena, California: Jet Propulsion Laboratory, California Institute of Technology, January 29, 1960); Nicholas A. Renzetti, " DSIF in the Ranger Project," Astronautics, September 196 1, pp. 3 5-37, 70.

9. Manfred Eimer, Albert R. Hibbs, and Robertson Stevens, Tracking the Moon Probes (JPL EP 701, Revised. Pasadena, California: Jet Propulsion Laboratory, California Institute of Technology, December 28, 1959), p. 9; Manfred Eimer and Y. Hiroshige, "Evaluation of Pioneer IV Orbit Determination Program," Seminar Proceedings: Tracking Programs and Orbit Determination, February 23-26, 1960 (Pasadena, California: Jet Propulsion Laboratory, California Institute of Technology, 1960), p. 43.

10. The Ran Project: Annual Report for 196](U) (JPL TR 32-24 1. Pasadena, California: Jet Propulsion Laboratory, California Institute of Technology, June 15, 1962), p. 383.

11. Nicholas A. Renzetti, ed., A History of the Deep Space Network from Inception to January 1, 1969 (JPL TR 32-1522. Pasadena, California: Jet Propulsion Laboratory, California Institute of Technology, September 1, 197 1), Volume 1.

12. JPL Interoffice Memo No. 81 from William Pickering to Senior Staff, Section Chiefs, and Section Managers, subject: "Laboratory Policy and Procedure for Space Flight Operations," November 9, 1960 (2-305); see also, Marshall Johnson, "Space Flight Operations," Astronautics Information: Systems Engineering in Space Exploration Seminar Proceedings, May 1-June 15, 1963 (Pasadena, California: Jet Propulsion Laboratory, California Institute of Technology, June 1, 1965), p. 39.

13. JPL Interoffice Memo No. 114 from William Pickering to Senior Staff, Section Chiefs, and Section Managers, subject: "Establishment of Space Flight Operations Section, " October 5, 1961 (2-299).

14. JPL Report from the Data Handling Committee to Brian Sparks, subject: "Interim Report from Data Handling Committee," June 22, 1961 (3-666); letter from Val Larsen to Abe Silverstein, subject. "Data Operations Command Facility Summary and Index, " October 17, 1961 (2-1460a).

15. NASA News Release, Homer E. Newell, "Remarks on the Dedication of the Space Flight Operations Facility, Jet Propulsion Laboratory, " May 14, 1964, p. 9 (3-170c).

16. See United States Congress, House, Committee on Science and Astronautics, Management and Operation of the Atlantic Missile Range, 86th Congress, 2nd Session, 1960.

17. Interview of James Burke by Cargill Hall, January 27, 1969, p. 4 (2-1391)

18. Letter from Richard E. Homer to Bernard Schriever, March 17, 1960 (2-2263); see also, Francis E. Jarrett, Jr., and Robert A. Lindemann, Historical Origins of NASA's Launch Operations Center to July 1, 1962 (Comment Edition, KHM-1. Cocoa Beach, Florida: John F. Kennedy Space Center, National Aeronautics and Space Administration, October 1964), pp. 66-67 (5-225); and paragraph 4.a of the LOD/AFMTC Agreement of July 17, 1961.

19. See Abe Silverstein's comments," Minutes of the Space Exploration Council Meeting, April 25-26, 1960, " p. 11 (2-1406).

20. JPL Interoffice Memo from Clarence Gates to All Concerned, subject: " Launch to Injection Instrumentation, " June 14, 1960, p. 2 (2-1414).

21. JPL Interoffice Memo from Joseph Koukol to Clifford Cummings, subject: "Comments of the Agena Tracking Panel Meeting at AMR, June 8, 1960," June 10, 1960, p. 1 (2-1039).

22. TWX from Sam Snyder, NASA Headquarters, to Kurt Debus, LOD, subject: "AMR Atlas Agena-B Launch Management," July 19, 1961, in NASA Agena Program Presentation, October 1, 1962, p. 3-18 (2-2269).

23. Range Use and Support Agreement Between the Launch Operations Directorate, George C. Marshall Space Flight Center, NASA, and the Air Force Missile Test Center, Air Force Systems Command USAF, at Patrick Air Force Base, Florida," signed by Kurt H. Debus, Director LOD, and L.1. Davis, Major General, Commander AFMTC, July 17, 1961 (2-23 74).


Ch 4 Notes  link to the previous page                         link to the next page  Ch 6 Notes

Chapter 4  link to the previous page        link to the next page  Chapter 6