SP-4306 Engines and Innovation: Lewis Laboratory and American Propulsion Technology

 

CHAPTER NINE

THE APOLLO ERA: OPPORTUNITIES AND DASHED HOPES

 

[177] Abe Silverstein returned to Lewis Laboratory optimistic that the success of the Apollo Program would be a major stepping stone in America's exploration of the cosmos. Trips to Mars and beyond would require new propulsion systems. Lewis Laboratory was poised to take the lead in NASA's program in electric propulsion. Silverstein also envisioned a major role for the laboratory in the proposed $1 billion program to develop nuclear rockets. If funded, Lewis would become the hub of NASA's nuclear propulsion program. NASA Administrator James E. Webb confirmed this commitment in a speech at a gala dinner at Cleveland's Auto and Aviation Museum hosted by the laboratory's loyal advocate, Frederick C. Crawford.1 Yet the promise of 1961 did not become reality, and eight years later Silverstein would leave NASA bereft of his earlier sense of optimism for either Lewis or the space agency as a whole.

Four years in Washington had not changed Silverstein's physical appearance or his basic management philosophy. His compact build, slightly rounded shoulders, baggy suit, and carefully groomed jet-black hair seemed no different than the day he had arrived from Langley Field in 1943. Part of the Langley nucleus that had carried NACA traditions from Hampton, Va., only his title had changed. He was now Dr. Silverstein. In 1958 Case Institute of Technology had granted him an honorary engineering degree.

Silverstein had seen the facilities of the Cleveland laboratory rise from the flat fields that skirted the edge of Rocky River ravine. He had participated in the transition from the piston engine to jet propulsion and, as Chief of Research, had guided the laboratory through the thicket of turbojet propulsion problems from the late 1940s through the early 1950s. Silverstein liked to have a personal sense of the projects under his supervision. During his short trips across the parking lot to the Engine Research Building for unannounced visits, he could take in the metallic glow of the Altitude Wind Tunnel, a facility with its own distinguished but tortuous history. There Silverstein and his staff had generated key data to redesign piston engine baffles so that cylinders would not overheat. After World War II, they had constructed small supersonic tunnels tied into the Altitude Wind Tunnel's air intake and exhaust system to explore the aerodynamics of ramjets.2

The overhead pipes that connected the tunnel with the laboratory's central air system symbolized the network of human connections that breathed life into its complicated facilities. These facilities embodied the creativity of Lewis personnel. While Silverstein was away, the tunnel's cavernous space had been transformed into a vacuum chamber used to test the instrumentation for Big Joe. After the Atlas-D rocket had lofted Big Joe into space, NASA knew the Mercury capsule would be safe to carry humans.3

 


[
178]

NASA Administrator James Webb envisioned Lewis as the hub of NASA's nuclear engine project.

NASA Administrator James Webb envisioned Lewis as the hub of NASA's nuclear engine project. Left to right: Abe Silverstein, Frederick C. Crawford, T Keith Glennan, James Webb. (Photograph Courtesy of Rebman Photo.)

 

 

Silverstein looked forward to working again at the laboratory level. He still had the rough edges of an engineer, more at ease with hardware than lobbying. At Headquarters, T Keith Glennan had found Silverstein articulate, but stubborn. In Glennan's view the NACA management style had suited the NACA, but "hardly fits our needs today when we have so many relationships with industry and other elements of the government."4 Silverstein "could be gruff, tactless, and impatient with mediocrity, hardly the way to win friends in Washington. His engineering colleagues, however, recognized his leadership ability. In a New York Times interview, a NASA colleague called Silverstein a genius - "not in terms of invention and discovery but in his breadth of comprehension of technical matters and his remarkable facility for getting down to the fundamentals in any field he tackles."5 With an intuitive feel for technical detail, he often became too involved when a project excited him.

Silverstein knew how to build effective engineering teams by carefully knitting together personalities and skills. He had a sure grasp of who moved best by being left alone and who had to be prodded to action. He could appreciate both ends of the research spectrum - not only the skill of the technicians in the machine shop, who could create a delicate instrument from an engineer's rough sketch, but also the analytic talent of the small group of research scientists at Lewis Laboratory.6 During the NACA years, engineering applications had dominated the activity of the [179] laboratory. Silverstein knew that NACA-trained NASA personnel were often looked down on by staff at the jet Propulsion Laboratory (JPL) and Marshall Space Flight Center and by the Vanguard team at Goddard Space Flight Center because of the emphasis on applications of NACA's work.7

Glennan had not fully appreciated the NACA research tradition. The Space Act required that NASA supervise industry contractors and manage space missions. The former college president, a Republican, favored spending most of NASA's budget on contracts with the aerospace industry and universities.8 He disapproved of the NACA policy of "farming out to industry only the repetitive and straight production items," saving the more challenging and creative projects for the staff.9 Although the "proper balance" between in-house research and industry contracts remained unresolved, as long as Glennan set a deliberate pace for space flight development, the integrity and autonomy of the former NACA laboratories seemed secure. After President Kennedy's announcement of the goal of a lunar landing, however, the tight schedules of the Apollo Program threatened to compromise the independence of the former NACA laboratories to do basic research. Under Webb's new administration, their role was no longer clear.

 

WHITHER LEWIS LABORATORY?

 

Even before Silverstein returned to Cleveland, the laboratory had begun to struggle with the implications of the new NASA organizational structure. Like the other former NACA laboratories, Langley and Ames, it was now referred to as a research center to distinguish it from the new development- and mission-oriented centers within NASA. It was clear that the NACA research ideal needed modification. In a speech to the staff the preceding year, Acting Director Eugene Manganiello revealed his anxiety about how the work of the center fit into the larger NASA picture. He summarized the history of Lewis to get a clearer sense of current practice and future goals. He pointed out that the amount of research in rockets and nuclear propulsion had remained small until 1951, when these areas began to grow. In 1957 Lewis Research Center had dramatically reduced research on air-breathing engines. By 1960 work on chemical rockets was at 35 percent; nuclear propulsion, 20 percent; and electric propulsion and power generation, 14 percent. Air-breathing engine research was at a mere 7 percent, and was mainly in support of military projects such as the B-70 airplane. The remaining 24 percent included basic heat transfer, fluid mechanics, radiation physics, and instrument and computing research."10

Manganiello considered the primary mission of Lewis Research Center to lie in three research areas: advanced propulsion, including chemical, nuclear, and electric rockets; power generation; and materials. He thought that the center should, as in the past, maintain a mixture of basic research, applied research, and what he called "specific research" related to development problems, such as the restart capability of the Centaur engine, the Saturn base heating problem, and the B-70 inlet control problem.11

In the NACA era, Lewis workers had responsibility only for the "research end of the business." Industry had handled development and production, and the military had been responsible for operations. Now, however, the situation was different. Space exploration required national resources of a magnitude that precluded maintaining the distinctions between research, development, and operations. "Our place as a research center is not as unique and sharply delineated as it was in the past." NASA had the responsibility not only for research, but also for the development, procurement, and operation of space vehicles. Manganiello recognized that some former NACA staff thought that research and development should be kept separate. "The [180] argument is that once the development camel's head is permitted in the tent it will inevitably take over completely and crowd research out of the shelter."12

Manganiello thought that the sharp distinction between research and development did not apply to current practice. "Applied research" and "advanced development" had become indistinguishable. The development centers - Marshall, JPL, and Goddard - did both. The research character of the former NACA laboratories, however, could be preserved if short-term development problems were avoided.

 

We at Lewis have no desire or intention to engage in the short-range activity or get into the problems of the construction of flight hardware. We are not particularly intrigued with the battle of the "O" rings but we are, on the other hand, intensely interested, in the advanced research and development end of the business, the nuclear rocket, electric propulsion and power generation, high-energy chemical rockets - their systems and components - where the scientific and engineering technology needs to be established before useful and reliable propulsion can be created. We cannot afford to permit our basic research or our conceptual research into new and better ideas to become submerged or deemphasized so that we have nothing left to grow on.13

 

Manganiello did not mention Lewis Research Center's contributions to the Mercury program because exclusive focus on these types of projects - albeit necessary and valuable in terms of the space race against the Russians - would leave the laboratory without research capital for the future. A research laboratory was a place where unconventional new technology like nuclear, electric, and high-energy chemical rockets could be incubated. This advanced work had to be nurtured while some of the laboratory's efforts could be diverted to assist in trouble-shooting when NASA needed their particular expertise.

Abe Silverstein could appreciate the ideas in Manganiello's talk. For almost four years in Washington, he had tasted the drama and excitement of large development projects. However, he was also aware that the quickened tempo of the Apollo Program now threatened the balance between research and development within NASA as a whole. Dryden had favored preserving the research orientation of the former NACA laboratories. He had placed them under the' Office of Advanced Research and Technology (OART) to protect them from the encroachments of the development and operations side of NASA. The integrity of research was a cornerstone of....

 

 


Eugene Manganiello, Acting Director after Sharp retired in 1960.

Eugene Manganiello, Acting Director after Sharp retired in 1960.

 

[181] .... Post-World War II technology planning. Vannevar Bush had regarded research, both basic and applied, as a government function. It was the nation's technical capital. In Science, the Endless Frontier he had warned, "Research will always suffer when put in competition with operations."14 Manned space flight now gobbled up almost all of NASA's budget. Just as Webb's new management structure had eased many former NACA staff out of Headquarters, the camel of development and missions had begun to push Lewis Research Center, along with the other two research centers, Ames and Langley, out of the mainstream of NASA.

Silverstein's return rekindled the debate between staff committed to basic research and those who favored a vigorous involvement in development and operations. Should the laboratory give up its pretensions as a research center and join the quickened pace of NASA centers more closely identified with development? Most of the laboratory's work was done in-house. Silverstein felt that the strength of the laboratory lay in the organic, self-sufficient engineering environment it provided. This environment provided the conditions for the growth of individual engineering expertise and the freedom to cultivate areas of research beyond the secure arena of existing technology. Within NASA as a whole, however, the concept of in-house research was disappearing. NASA began to contract out support services, such as janitorial, grounds-keeping, and security. Eventually, support service contracting would include basic technology research. A windfall for the aerospace industry, this policy threatened the mission of the former NACA laboratories to perform basic and applied research. In 1961, 77 percent of NASA's employees were contractors. By 1964, contractors would command 92 percent of NASA's employment total.15

Silverstein had not decided what direction the laboratory should take when he returned to Lewis. Increased involvement with the development and mission side of NASA would mean that the laboratory would have to learn to deal with contractors. He also knew the danger of letting the camel of development into the research tent. How would he protect basic research? If Silverstein protected the research character of the laboratory by refusing to participate in large-scale development projects, he was aware that he needed more staff with graduate educations. Lewis's group of theoreticians was small. The majority of staff had degrees in mechanical engineering. When three members of the Applied Mechanics Group - Simon Ostrach, Stephen Maslen, and Harold Mirels - called on him to discuss the laboratory's future, Silverstein took heed. Ostrach recalled Silverstein's surprise at their audacity in proposing that he transform Lewis into a laboratory for basic research. "They sort of put us off in a corner and they thought that they should have bright guys around, but Abe was flabbergasted that we wanted to have something to say about the direction that the lab was going."16

At a meeting of division chiefs Silverstein presented the idea of the Applied Mechanics Group to model Lewis on AT&T's Bell Laboratories in Murray Hill, N.J. He wanted to hire only Ph.Ds. Although the general consensus was favorable, Bruce Lundin strongly opposed it. He had spent most of his career testing full-scale engines. To Lundin, Dryden's often repeated admonition that the NACA laboratories needed protection from the development and mission side of NASA had become a shibboleth. He argued that they were "really just a bunch of testing engineers." To transform the laboratory by adding a group of high-powered scientists would take ten years. Lundin thought that NASA needed the laboratory's technical competence. He suggested a compromise: divide the laboratory into two distinct parts, research and development. Silverstein, as director, could "provide the balance between the two and can protect the one from the other."17 Lundin's plan seemed to offer the best of both worlds. With the backing of the other division [182] chiefs, Silverstein reorganized the laboratory. He put Lundin in charge of development and John Evvard in charge of research.

The laboratory immediately began to experience the effect of strong Congressional support for the space program during the early Apollo years. Silverstein announced that Lewis Research Center would increase its payroll from $23.5 million to $28 million. He was authorized to hire 615 new staff, a major expansion of personnel. In addition, a large Developmental Engineering Building to house 1100 engineers was planned.

The location of this large office building outside the main gates reflected Silverstein's commitment to keep research physically separate and untarnished by the contracting side of NASA. Lundin's staff, many newly hired under NASA's more generous salaries, worked in the Developmental Engineering Building, with its separate cafeteria and security. They would learn the new realities of dealing with industry as a customer rather than a provider of technology. At first, the research side of the center was shielded from responsibilities connected with contracts. People who worked under Evvard still came through the Main Gate in the morning to work in their laboratories much as they had during the NACA era. However, even the research side of the laboratory began to be affected by the pressure of increased NASA-wide contracting.

 

TO THE MOON: DIRECT ASCENT OR RENDEZVOUS?

 

In November 1961, when Silverstein returned, it appeared that Lewis Laboratory would become an important center for nuclear rocket propulsion. With a nuclear stage proposed as the key to providing sufficient power to reach the Moon directly, Lewis Research Center was positioning itself to play a leading role in NASA. The Cleveland Plain Dealer reported that Lewis Research Center looked forward to "major participation in every manned space effort to follow the current Mercury Program and most future unmanned space exploits as well."18 The Kennedy Administration was reluctant to approve costly expenditures for large technology development programs unless they could be justified in terms of specific applications.19 The costs of the nuclear programs were astronomical, but tied to future missions beginning with Apollo; NASA counted on obtaining generous funding for nuclear rocket research.

The ambitious lunar mission called for a rocket that could provide enormous thrust, nothing less than von Braun's huge Saturn or an even larger rocket called Nova, first proposed by the Saturn Vehicle Evaluation Committee, chaired by Silverstein, in 1959. No consensus was ever reached on Nova's configuration, but one proposal considered a first stage powered by conventional fuel, a liquid hydrogen second stage, and nuclear upper stages.20 The nuclear stages could give this behemoth the necessary thrust for direct ascent to the Moon, which involved a scenario reminiscent of the science fiction of Jules Verne. A rocket consisting of several stages would be fired directly at the Moon. It would need to brake against the Moon's gravity to land. For the return journey to Earth, a final stage would blast off from the Moon's surface. The first study of launch vehicle requirements for the lunar mission, chaired by William A. Fleming, formerly of Lewis, supported the direct ascent concept.21 Projected costs and time for the development of a rocket of the required magnitude made some NASA planners question the feasibility of direct ascent.

Within NASA there were also strong partisans of a rendezvous method to reach the Moon in two stages via a space platform placed in orbit around Earth. von Braun and the Marshall group favored what came to be called the Earth-orbit rendezvous method. Two packages consisting of modules for the assembly of the platform and lunar vehicle could be launched separately by two [183] Saturn rockets to rendezvous in Earth orbit. The platform would then function as a launching pad to send the lunar vehicle to the Moon and back. The attraction of the Earth-orbit rendezvous method was that instead of a single large rocket, which existed only in the imagination of NASA's rocket designers, the plan depended on the smaller Saturn rocket already under development at Marshall Space Flight Center. By June 1961, after a NASA study chaired by Bruce Lundin, opinion began to shift in favor of the rendezvous concept, with Earth-orbit rendezvous the "clear preference" among members of the evaluation committee. Nova development, however, was still under serious consideration.22

A group at Langley Research Center favored a third approach-lunar-orbit rendezvous (LOR) - which at first seemed hopelessly complicated. It required three spacecraft: a command module occupied by the three astronauts, a service module for the propulsion and guidance systems, and a lunar excursion vehicle. These three vehicles would be fired into lunar orbit by a single expendable three-stage rocket. Once in lunar orbit, the astronauts would park the command and service modules in orbit. Two of the astronauts would don space suits and clamber into the excursion vehicle for their trip to the Moon's surface. Later they would rendezvous with the command module for the return trip to Earth, leaving the excursion vehicle behind. The advantages of this method were its lower fuel and weight requirements.

Silverstein favored direct ascent. He objected that the engineering required for the docking of two vehicles in space involved greater complexity and more risk to the astronauts should a component fail.23 Although Silverstein returned to Lewis before NASA planners had hammered out the final decision, no doubt it came as a disappointment. Webb's announcement of the selection of the lunar-orbit rendezvous method in July 1962 ended all hopes for the development of Nova and took some of the urgency away from Lewis's nuclear rocket program. Nuclear Power might be required for distant flights to Mars and beyond, but clearly, through the 1960s, NASA's energies would be directed to landing humans on the Moon.

Lewis staff wanted to salvage something from the lunar rendezvous decision. Because of the laboratory's expertise in propulsion, Lewis was the logical choice to design and monitor the development of the second vehicle, called the Lunar Excursion Module (LEM), needed to take the astronauts from the command module to the surface of the Moon. As early as December 1959, Lewis staff had set its sights on the development of a "lunar soft-landing vehicle." E. W. Conrad and Carl F. Schueller took up the idea at Headquarters with Cortright in 1959, long before NASA had decided on the details of the Apollo mission. Although no formal agreement was reached, Cortright gave them the go-ahead to begin to work up specifications. By 1960, the Analysis Branch of the Propulsion Systems Division was at work on two types of vehicles for manned lunar-landing missions. The first was a small-scale vehicle for lunar exploration to carry two or three astronauts, with a return capsule weight of about 10,000 pounds. A second larger vehicle to deliver 50,000 pounds of material to a large-scale lunar base was also designed. However, with Silverstein back at the helm of Lewis Research Center, this project had to be dropped. The develop the LEM would have required working with von Braues group, in charge of the Saturn V. Although there may have been other considerations, in Bruce Lundin's opinion, Brainerd Holmes opposed creating this potentially abrasive situation. He was "not going to have Silverstein throw sand into his machine."24

 

 

THE TANGLED HISTORY OF NUCLEAR PROPULSION

 

[184] The Lewis staff were old hands at dealing with the opportunities and disappointments associated with the development of nuclear propulsion. Interest at the laboratory in nuclear propulsion had developed steadily from the late 1940s under the leadership of Benjamin Pinkel and his branch chief, Eugene Manganiello. In 1948 a prestigious group at the Massachusetts Institute of Technology issued the secret Lexington Report, which concluded that development of nuclear rockets and ramjets presented nearly insuperable technical problems. Aircraft nuclear propulsion, however, was feasible, although development might take up to 15 years. To the American propulsion community, led by Colonel Donald Keirn and D. R. Shoults of General Electric (both prime players in the secret drama to import the Whittle engine), this was sweet music. They thought an engine powered by nuclear energy could create a propulsion "breakthrough" comparable to the turbojet.25

Silverstein envisioned increased cooperation with the Atomic Energy Commission and a strong role for the Cleveland laboratory in NEPA (Nuclear Energy Propulsion for Aircraft). In 1949 the laboratory acquired a cyclotron for basic research in materials.

Carving out a role for Lewis in aircraft nuclear propulsion proved an agonizingly slow process. Planning for a nuclear reactor began in 1954, and in 1955 Congress authorized its construction. After a survey of 16 locations in Ohio and Pennsylvania, Lewis leased 500 acres of land near Sandusky, Ohio, about 50 miles west of Cleveland. Originally known as the Plum Brook Ordnance Works, it had served as an explosives factory and storage area during World War II. In 1956 the Atomic Energy Commission Safeguard Committee approved the design for the reactor. At the time of the ground breaking for the new facility, aircraft nuclear propulsion appeared to be the propulsion frontier. Silverstein called it "the "shining hope' for increasing the range of aircraft at high speeds and for increasing aircraft ranges to values unobtainable with conventional or special chemical fuels."26 The laboratory hoped to contribute fundamental studies on the effects of radiation on materials. Only gradually did it become apparent that the nuclear airplane had become a technical dead end. Its detractors called it a shitepoke - an enormous skinny bird, hardly fit for eating or for flying. Development by Pratt & Whitney and General Electric over 15 years cost the nation $880 million, but as late as 1960, how it would benefit the nation's defense remained unclear. The weight of the shielding for the reactor, as well as a new awareness of environmental considerations, led to the national program's demise in 1961.27

Ironically, as interest in the nuclear airplane waned in the late 1950s, enthusiasm for a nuclear rocket waxed. In 1955 Robert W Bussard of the Oak Ridge National Laboratory questioned the conclusions of the Lexington Report. He argued that the key to nuclear rocket development was temperature-resistant materials. His advocacy convinced the Air Force and the Atomic Energy Commission (AEC) to set up a joint program at Los Alamos Scientific Laboratory to develop Rover, a nuclear rocket intended to be launched from the ground.28 Because hydrogen was the preferred propellant for nuclear rockets, the growing expertise of the Lewis staff in handling this fuel put them in high demand. They assisted in the design and testing of the KIWI series of experimental reactors, managed jointly by the AEC and NASA, at Jackass Flats, Nev.29 Frank Rom, Chief of the Nuclear Propulsion Concepts Branch at Lewis, served as the laboratory's chief spokesman for its programs in materials, fuel element research, and hydrogen heat transfer.

The AEC, however, controlled the nuclear field through tough licensing requirements. It regarded NASA as an interloper in the nuclear field. T. Keith Glennan lamented that the Plum Brook reactor was "proposed and accepted at a time when the aircraft nuclear propulsion work [185] was at its white hot heat."30 Glennan stoutly defended the quality of the reactor's design, but whether Lewis could land a role in nuclear rocket development was in doubt. Senator Clinton B. Anderson of New Mexico was determined to see that funding went to the AEC's laboratory at Los Alamos. When NASA and the AEC set up a joint office called the Space Nuclear Propulsion Office (SNPO) in 1960, the two federal agencies began to enjoy a smoother relationship. The future of Lewis's programs in nuclear propulsion looked brighter. Harold Finger took charge of both the SNPO Office and NASA's Nuclear Systems Division.31

Despite the intensity of research on the part of NASA and AEC staff, the use of Rover in NASA's stable of launch vehicles was remote, Glennan asked the obvious question: "Just where one would launch such a beast with its ever present possibility of a catastrophic explosion resulting in the spreading of radioactive materials over the landscape is not clear."32 In 1961 the Rover project was renamed NERVA (Nuclear Engine for Rocket Vehicle Applications).33 After Congress authorized a major three-year $40 million building program for Plum Brook in the fall of 1962, Silverstein set up a design group to plan a $15 million facility for testing a nuclear rocket engine at Plum Brook. Although a nuclear upper stage for the Saturn launch vehicle seemed increasingly unlikely, the feasibility of a nuclear rocket for post-Apollo missions had to be demonstrated.

 

RETURN TRIP T0 MARS

 

Lewis planners believed that landing on the Moon was only the first step in space exploration. NASA's next destination would be Mars or another planet. Chemical rockets like Saturn, or a nuclear rocket, could reach a near planet, but longer trips required different types of rocket systems. The advantage of an electric rocket was its low propellant consumption and continuous long-term operation. The group formed by Wolfgang Moeckel after the success of "From Mach 4 to Infinity" began studies of electric propulsion and space power systems that involved plasma physics and magnetogasdynamic and thermionic systems. They also investigated the possible applications of controlled nuclear fusion for space propulsion. The group explored the possibilities of using high-intensity, large-volume electromagnets with the lowest possible mass for power generation. In 1958 they initiated a small program on plasma heating to complement the studies of magnetic fields.34

Initially, the electric propulsion systems under investigation consisted of two major components, the electric power generator and the thrust generator. The electric power generator - either nuclear or solar-converted energy into electric power. The thrust generator used this power to accelerate the propellant out the back end in the form of thrust. The 1957 study of a nuclear turboelectric power plant indicated that sodium had the potential to make a good heat transfer fluid. By passing the fluid from the fission reactor through a neutron shield, a heat exchanger, and back to the reactor, the crew could be protected from radiation. The generator would provide a power output of 20, 000 kilowatts, with 11,000 kilowatts of power in the form of thrust. A round trip to Mars would require a vehicle weighing 350,000 pounds. In addition to the structural weight of the vehicle, the eight-man crew would need 50,000 pounds of equipment and a 40,000-pound auxiliary rocket for landing part of the crew on Mars. 35 Admittedly fanciful, this power system was the basis of more sophisticated design studies and hardware development related to nuclear-turboelectric systems.

The enthusiasm and expertise of the group grew under the leadership of Howard Childs, William Mickelson, and Wolfgang Moeckel. Their work embodied Dryden's vision of advanced....

 


[
186]

Nuclear Rocket.

Nuclear Rocket.

 

....technology research as distinguished from development. Moeckel explained in a talk at Head quarters in 1961:

 

The approach taken by the Lewis Research Center in its electric propulsion program is to undertake studies of problems common to all systems. This approach is in contrast to that of development organizations, which must generally concentrate their effort on perfecting a particular system, based on current state-of-the-art. There are, of course, many areas of overlap, but in general the Lewis effort is directed toward obtaining the fund of knowledge and the new approaches required to design specific systems for various applications, rather than to design or produce those systems directly.36

 

The group worked with the AEC to develop the nuclear-turboelectric system, SNAP-8, intended to provide 35 kilowatts of on-board power in space. The first solar project, Sunflower, a joint effort with Thompson-Ramo-Wooldridge (TRW), yielded basic knowledge of dynamic solar power systems for spacecraft. The generous budgets for construction of facilities during the Apollo era enabled the staff to build a large vacuum tank for electric propulsion studies. Because of Lewis's new facilities and the large number of staff involved, in 1961, shortly after Silverstein's return, NASA announced that Marshall Space Flight Center's electric propulsion program would be transferred to Lewis.

 


[
187]

Water filled quadrants surrounding the Plum Brook nuclear reactor, 1961.

Water filled quadrants surrounding the Plum Brook nuclear reactor, 1961.

 


[
188]

Diagram of electric space vehicle with nuclear power source for a round trip to Mars.

Diagram of electric space vehicle with nuclear power source for a round trip to Mars.

 

 

The laboratory environment supported the interplay between theory and hardware necessary to foster technical creativity. Technical support and relative freedom to pursue promising ideas regardless of immediate application allowed Harold R. Kaufman to design a more efficient and simpler electric rocket than Stuhlinger's earlier cesium thruster.37 Kaufman was a member of the Mission Analysis Branch. He and a colleague, Wilbur Dobson, knew of Moeckel's interest in Stuhlinger's work but were not members of his division. Kaufman took a different approach. His invention, the electron-bombardment ion thruster, used mercury vapor as a propellant. The vapor flowed into an ionizer chamber; the mercury ions were then propelled toward a screen grid, which accelerated them to produce thrust.

In 1964 the SERT I (Space Electric Rocket Test) tested both Stuhlinger's cesium engine, built by Hughes Research Laboratories, and Kaufman's electron-bombardment thruster, designed and built at Lewis. Kaufman's approach proved superior. It also demonstrated that a stream of ions from a thruster in space could be neutralized to avoid the buildup of a space charge that would shut down the thruster. This established that ion thrusters could be made to work in space. The development of the ion-bombardment thruster culminated in 1970 in the successful launch of SERT II after ten years of painstaking development of Kaufman's design. SERT II, however, reached fruition at a time when the ability of NASA to support advanced technology was flagging.38

 

CENTAUR: ABE'S BABY

 

Centaur, the third program that came to prominence during the Apollo era, was a legacy of the laboratory's work on liquid hydrogen in the 1950s. Centaur was a development program urgently needed for the success of a lunar landing. It was originally funded through the Advanced Research Projects Agency (ARPA) and assigned to Marshall Space Flight Center after President Eisenhower insisted that all space programs without direct military applications be taken over by NASA. After NASA acquired the von Braun team, Centaur was designated the launch vehicle for Surveyor, the unmanned lunar lander program. It was managed under NASA's Office of Launch Vehicle Programs, clearly the development and operations side of NASA. Centaur came to Lewis by default. It was an unwanted step-child at a time when all the efforts of von Braun's rocket group were focused on the development of Saturn.

In September 1962 Edgar Cortright, then Deputy Director of the Office of Space Sciences at NASA Headquarters, remembered the keen interest in liquid hydrogen at Lewis. He asked Silverstein to come to Headquarters where he brandished a letter that his boss, Homer D. Newell, [189] had received from Wernher von Braun. von Braun thought that NASA should cancel the Centaur Program. von Braun apparently was not yet convinced of the feasibility of liquid hydrogen as a fuel, despite his capitulation to Silverstein during the Saturn Evaluation Committee meetings in December 1959. He was also concerned about Centaur's structure. According to John Sloop:

 

The von Braun team simply did not trust Karl Bossart's thin-wall, pressure-stabilized tanks first used by General Dynamics for Atlas and then for Centaur. Centaur had only a thin, common wall dividing the fuel and oxidant tanks, and at one time von Braun came to Washington to push a design change that separated the two tanks, a change that would delay the stage development. His request was denied by Robert Seamans. I believe that these structure concerns outweighed von Braun's concerns about using liquid hydrogen.39

 

von Braun recommended that the Saturn C-1/Agena D launch vehicle replace Centaur. The jet Propulsion Laboratory (JPL), in charge of the Surveyor Program since 1960, concurred. von Braun convinced JPL that Centaur could not be developed in time to provide the necessary knowledge of the Moon's surface prior to the Apollo landing. In a letter to Newell, Brian Sparks, Deputy Director of JPL, cited "the deplorable situation in the current Centaur program" in support of cancellation. The development of a new rocket to use liquid hydrogen technology required too much time. To compete with the "doggedly determined effort of the Soviets," the existing technology of kerosene-based fuels had a greater chance of success. He wrote:

 

The Office of Manned Space Flight has said that data from a soft-landed Surveyor in 1964 or 1965 would be an invaluable aid in Apollo design, and high-quality reconnaissance data from a lunar Orbiter in 1966 is desirable for Apollo flight planning. These objectives could be achieved on a C-1/Agena D but not on Centaur in this time scale.40

 

Problems with NASA's contractors, General Dynamics and Pratt & Whitney, plagued the development of Centaur. However, the supervision of these contractors by the von Braun team had also played a role in its recent failures on the launch pad. On May 8, 1962, 54 seconds after liftoff, the Centaur portion of an Atlas-Centaur launch vehicle had exploded. The investigation in to this mishap revealed that an internal NASA report had predicted this failure: the insulation panels could not withstand the anticipated pressure loads. The accident investigators concluded that, although General Dynamics was responsible for the defective design, Marshall Space Flight Center's supervision had been neither prompt nor adequate. In addition, the investigation censured Marshall's supervision of Pratt & Whitney's development of the RL-10 liquid-hydrogen engine. Three explosions of engines on Pratt & Whitney's test stands at its Research and Development Center in Florida had resulted in both delay and $1.2 million worth of damage, The investigation concluded that these were preventable accidents. The company had failed to install standard safety devices that Marshall engineers, with extensive experience in the hazards of rocket engine development, should have insisted on.41

Cortright informed Silverstein that the relationship between Marshall and General Dynamics had reached an impasse. Both were aware that project management required a healthy give-and-take. Technical differences had created an adversarial relationship between General Dynamics and Marshall.42 Marshall had accepted Centaur, but von Braun, never comfortable with liquid hydrogen or the Centaur tank design, did not assign his best engineers to the project. [190] In Silverstein's view, "They did not believe in Centaur because it was not really theirs. Centaur was not a program that they had initiated.43

In September 1962 the choice at Headquarters appeared to be either to drop the Centaur program entirely or find another NASA center willing to take over its management. When Edgar Cortright asked Silverstein if Lewis wanted the Centaur program, Silverstein' needed no arm-twisting. He believed that liquid hydrogen would provide the key to a successful lunar landing. The transfer of launch-base operations from Kurt Debus's group to the Goddard Field Projects Branch at Cape Canaveral under Robert Gray completed the decoupling of Centaur from Marshall.

When Silverstein announced to his division heads that NASA had assigned Centaur to Lewis, the response was lukewarm.44 Although Lewis had tasted project management when NASA created the Nuclear Engines Project Office and the Sunflower program, these were slow-paced programs. The management of Centaur posed development problems that needed immediate solutions. In October, plane loads of boxes containing books, drawings, spare parts lists, unsatisfied change orders, and technical directions were sent to Lewis from Marshall. For men and women accustomed to working in a research environment consisting of blackboards, glistening hardware, and the roar of facilities, the silent mountain of government paper produced dismay. What did they know about the legal tangle of government contracts?45

Centaur forced its new project managers to leave the comfortable campus atmosphere of the research laboratory to deal directly with industry. Silverstein picked a technical team to tackle the Centaur Program that matched the Lewis administrative talent he had left in Washington. David J. Gabriel, a Lakewood native with a mechanical engineering degree from the University of Akron, shouldered the responsibility of project manager. Cary Nettles, Russel Dunbar, Ed Jonash, Jack Brun, and John Quitter, seasoned Lewis staff, formed the technical core of the project. Although Centaur fell within Bruce Lundin's administrative bailiwick, Silverstein himself was never far from what was going on in the Centaur project.

Silverstein called on Lewis's small law department to bring order out of chaos. Neil Hosenball, the laboratory's chief legal counsel, asked Len Perry, a vibrant attorney with boundless energy and unusual powers of persuasion, to assume responsibility for the contractual aspects of the Centaur Project. He hired Harlan Simon, an Ohio State Law School graduate in private practice, to assist him. They attended all the Centaur team's technical meetings with General...

 

 


Harold R. Kaufman with his electron bombardment ion thruster.

Harold R. Kaufman with his electron bombardment ion thruster.

 


[
191]

Pratt & Whitney's RL-10 hydrogen-oxygen engine is readied for testing, June 1963.

Pratt & Whitney's RL-10 hydrogen-oxygen engine is readied for testing, June 1963.

 

[192] ....Dynamics to sort out the objectives of the program and where it had gone wrong. Simon recalled the intensity of Silverstein's involvement.

 

It was Abe's baby, Abe's project. He made a promise not only to the world, but to himself. He selected the people. He is the one who chose the people for these various positions and he chose them, obviously, correctly. He was informed of every decision that was made of any significance. He participated in the decision making process. He argued with them. They argued with him. It was both dedication to the program and fear of Abe that made this program successful. You just didn't dare let him down. Otherwise, I don't know what would have happened. You almost felt as though you would have been disintegrated, if you ever let Dr. Abe down.46

 

The lawyers had a formidable challenge: to restructure the business arrangement with General Dynamics to permit flexibility and mutual trust. They had to devise a contract that, above all, would not slow down the pace of development.

The original contract with General Dynamics was a not a firm fixed-price contract, but a cost-plus-fixed-fee contract of rather loose construction. It was not geared to the development of new technology under the pressure of a tight deadline. Any technical change required not only ponderous paperwork, but worse, an increase in the costs of the vehicle even before the final design had been established. A liquid hydrogen rocket required a leap into the unknown. It was a leap constrained by a tight schedule.

Hosenball, Perry, and Simon devised two types of contracts to cover the requirements of technology development. The first was the "hardware contract," which took "all the knowns and entered into a contract which had very definite boundaries in dollars and time." To cover the unknowns - risks of development, support for unexpected failure - they devised a "management and engineering contract."47 If a rocket exploded on the launch pad, another test could be done without waiting months for the necessary government paperwork to be completed. NASA could guarantee a profit to the company regardless of the outcome. This unconventional approach to government contracting could have left them open to criticism, but they were betting on success. The Centaur team worked at a frenzied pace, commuting to California to oversee work by General Dynamics and to Cape Canaveral to prepare for launches. Their labor was crowned with a string of successful launches, culminating in the flawless Surveyor mission in May 1966.

 

 


Centaur stage mated to Atlas booster at Plum Brook in the Dynamics Research Testing Stand, February 1964.

Centaur stage mated to Atlas booster at Plum Brook in the Dynamics Research Testing Stand, February 1964.

 

 

[193] Silverstein's commitment to liquid hydrogen was vindicated, and a similar success with liquid hydrogen in the upper stages of the Saturn V proved von Braun's pessimistic view of this tricky fuel unjustified. He autographed a picture of the launch of Apollo 4, on November 9, 1967: "To Abe Silverstein whose pioneering work in liquid hydrogen technology paved the way to today's success."48

What explains the intensity of this commitment? Why was the Centaur program so successful? This extraordinary combination of energy, teamwork, and leadership cannot be explained simply as the result of the ample funding that NACA received during the Apollo era. Nor can the achievement of the Centaur program be dismissed as an attempt by Lewis staff to demonstrate that they could succeed where Marshall had failed. Rather, the explanation for the achievement of the Centaur program lies within Lewis's institutional fabric, Unlike Marshall, Lewis was a research laboratory, The NACA research tradition had fostered the commitment to liquid hydrogen through the 1950s as Lewis researchers gained expertise in the problems of this complicated and dangerous technology. Silverstein epitomized the breed of technical leader that Arthur Squires has described as a "maestro of technology" - someone who is thoroughly familiar with the technology, who knows his staff well enough to put together the right team for a specific task, and who protects them from the aggravations of unnecessary bureaucracy.49 He inspired by example. His staff stayed through the night to untangle a technical problem because Silverstein did the same.

Beyond Silverstein were the extensive resources of a research laboratory. Lewis researchers faced a variety of problems in making liquid hydrogen usable as a rocket fuel. A research group in materials headed by Merv Ault studied the degrading effects of liquid hydrogen on various metals. Other Lewis staff examined how liquid hydrogen would behave in a weightless environment. They investigated pumps, turbines, and other components of the rocket engine. Testing in the 85-foot drop tower, and later in the zero gravity facility, provided data useful in modifying designs. Centaur staff had Lewis expertise in heat transfer to call on, as well as an extremely skilled group of technicians to build whatever part might be necessary to tease a recalcitrant rocket engine to life.

At the same time that Lewis inherited Centaur from Marshall, it acquired management of the Agena project. Agena was an upper-stage booster rocket usually coupled with a Thor or an Atlas rocket to launch spacecraft for NASA's planetary program. Unlike Centaur, Agena did not have the constant attention of Silverstein. Seymour C. Himmel took charge as project manager, flanked by his deputies C. C. Conger and E. F Baehr. Technical branches were headed by E. H. Davidson, M. Weston, and H. W. Plohr. They were all "NACA types" who had been at Lewis since the 1940s. They had never worked on launch vehicles before, but their NACA backgrounds gave them the flexibility, management skills, and technical expertise they needed to organize the staff (supplemented with former military contractors for intercontinental ballistic missile programs) and to build relationships with contractors, the Air Force, and centers like Goddard and JPL in charge of the spacecraft. Thirteen months after taking over the Agena project, the team launched its first two spacecraft, Echo II and Ranger VI. Between 1962 and 1968, when Silverstein assigned the Agena's payloads to the Centaur rocket, they launched 28 missions, including several Rangers, Mariner Mars '64 and Mariner Venus '67. The spacecraft launched by Agena sent back some of the first Pictures of Mars and Venus.50

The Centaur and Agena Programs were the glamour programs of the laboratory. Jobs in launch vehicles had visibility, mobility, and drama. The NACA-trained researcher looked on the....

 


[
194]

Photograph of launch of Apollo 4 with Wernher von Braun's message

Photograph of launch of Apollo 4 with Wernher von Braun's message "To Dr. Abe Silverstein whose pioneering work in liquid hydrogen technology paved the way to today's success."

 

 

[195] ....swashbuckling young engineering graduate recruited for project management with a mixture of pride and envy. Although research had more ample funding during the Apollo era than ever before, basic research lost status. The preparation of a research report, once the most respected activity of the laboratory, no longer stimulated the same intense commitment. The roar of a rocket at liftoff had supplanted the polite applause of engineering peers elicited by a well-presented paper. Some NACA-trained researchers like Harry W. Mergler, Simon Ostrach, Eli Reshotko, and Arthur Hansen drifted off to positions in academia. Basic work in materials, heat transfer, and tribology, the science of fuels and lubricants, quietly continued.

 

CASE INSTITUTE OF TECHNOLOGY AND THE NASA BANDWAGON

 

If the transformation of the NACA into NASA was a mixed blessing for former NACA laboratories like Lewis Research Center, universities stood to benefit from the sudden availability of large NASA grants. The NACA sponsored some university research through the 1940s and 1950s, but the amount of university grants had remained small because NACA laboratories preferred to do their research in-house.

The possibility of ample government funding infused Case Institute of Technology with new enthusiasm for aerospace engineering and a greater interest in cultivating connections with Lewis Research Center. Both Glennan, who returned to Case's presidency at the close of the Eisenhower Administration, and his successor, James Webb, supported strong NASA-university connections. In addition to specific NASA-sponsored research, NASA's charter stipulated that one of the agency's goals was to promote science education.

The first Case Institute of Technology proposal for an Interdisciplinary Materials Research Program in July 1960 described how its Cleveland location benefited Lewis Research Center. Case Institute's faculty had provided NASA with necessary additional training in materials in a recent in-house graduate course offered at Lewis. However, the former strains in the Case-NACA relationship may have influenced the proposal's rejection in favor of Rensselaer Polytechnic Institute.51

When Glennan returned to the presidency of Case Institute, he carefully laid the ground-work for an improvement in relations. In October 1962, he personally paid a call on Hugh Dryden and handed him a summary of the proposals awaiting action by NASA evaluators. In a memo to John Hrones, Vice President for Academic Affairs, Glennan revealed that the Case faculty thought it had been treated unfairly by NASA in the past: "I think that Dr. Dryden has protected the integrity of that memorandum which was rather critical of actions taken by one or more individuals in the NASA organization in the review of our proposals."52 Case faculty claimed that its proposals were still being turned down for reasons that had more to do with personalities than technical quality in 1963, but by October 1964 the situation had improved dramatically. Case was doing quite well. Its acceptance rate for the previous 12 months had reached 100 percent.53

In 1963 Case caught a whiff of NASA's plans for a high-tech facility. It strongly backed the city of Parma's bid for the site for a NASA Electronics Center, Not surprisingly, given the strong ties of the Kennedy Administration to Massachusetts, the city lost out to Cambridge.54 However, realistically, the Cleveland area had neither a cadre of Ph.D.s trained in electrical engineering nor strong industry support. Politics precluded two NASA facilities in the same city. Nevertheless, Webb saw a role for Case Institute of Technology in the government, industry, university partnership he envisioned. He hoped that Case would assist NASA in ferreting out Cleveland area industries that could use NASA-generated technology. He believed that space technology, turned to profits by industry, would mean a stronger American economy.

[196] However, it was one thing to formulate a policy, another to implement it. With the exception of TRW and Eaton Industries, there was little high-tech industry in Cleveland that could benefit from aerospace technology. Cleveland's economy, beginning to founder, depended on heavy industries like steel. It was increasingly clear that the role that Lewis had played in stimulating innovation among the engine companies during the NACA years would be difficult to recast for the utilization of space technology, despite the good intentions of Case and the Lewis Technology Utilization Program. A conference sponsored by Lewis to discuss new technology available to the electric power industry as a consequence of the space program created interest but no real commitment to innovation. Lewis would discover in work with the automobile industry in the 1970s the difficulties of pushing innovation on a reluctant industry.55

Although Case's involvement in the technology utilization program was minimal, it had better success with an ambitious proposal for a Space Engineering Research Laboratory, submitted in January 1965, The proposal for $2,580,000 in construction funds emphasized Case's strong programs in space-related disciplines: digital systems engineering; control of complex systems; bio and medical engineering; engineering synthesis; fluid, thermal, and aerospace sciences mechanics; and plasma dynamics. It proposed concentrating some of these programs in one building. The proposal indicated that the relationship with Lewis was beginning to flourish, with new staff recruited from Lewis and over 100 Lewis staff currently in advanced degree programs, 17 of whom were Ph.D. candidates. When the new NASA laboratory was dedicated in 1969, it was appropriately named the Glennan Space Engineering Building to honor Glennan's dedication to Case, NASA, and the AEC.56

In 1966 NASA turned renewed attention to aeronautical research, neglected after Sputnik. At Lewis, although the major effort remained in space-related activities, several new projects were initiated. The return to aeronautics among NACA-trained staff meant a return to the intriguing problems of compressors and turbines. The phenomenal growth of air travel had caused airport congestion, overcrowding, noise, and chemical pollution. The development of quieter engines, as well as aircraft that could take off and land on short runways, offered the promise of new technical challenges. Lewis also took part in plans for a supersonic transport airplane to compete with the Concorde under development by the British and the French. However, after a hiatus of nine years, could the staff return to a position at the cutting edge of aircraft propulsion technology? Lewis's facilities for propulsion research were no longer unique. Not only had the aircraft engine industries developed their own facilities, but the Air Force now had its state-of-the-art wind tunnels at the Arnold Engineering Development Center in Tullahoma, Tenn.

Lewis had eliminated work on air-breathing engines when it shifted into space propulsion. Researchers at Ames and Langley never abandoned their work in aeronautics and did few space-related projects in the 1960s. Lewis's space work, particularly its management of Centaur and Agena, had swelled its budget in the 1960s, although a large portion of these funds went to contractors. In 1964 Lewis received $299.9 million for research and development, compared to $78.1 million and $40.3 million, respectively, for Langley and Ames. For construction of facilities, most of which were at Plum Brook, Lewis received $26.5 million, compared to $9.0 million for Langley and $11.9 million for Ames. In 1965 Lewis's research and development budget continued to rise. It received a peak of $323.2 million, compared to $106.6 for Langley and $54.2 for Ames. Funding for construction of facilities again gave Lewis a substantial budget in 1965: $18.6 million for Lewis, $7.2 for Langley, $13.6 for Ames. 57 However, the handwriting was on the wall. The [197] Vietnam War and the War on Poverty cut into NASA's budget. In 1966, as the Agena project was being phased down, Lewis's research and development budget declined by $733 million. The next year, it dropped by another $87.2 million as NASA experienced its first major cuts. The Hough riots in downtown Cleveland revealed the bitter incongruity of the expensive space race and the problems of the inner city. Staff cuts began to bite into the morale of the laboratory.

In 1969, when Apollo 11 astronaut Neil Armstrong stepped onto the pocked surface of the Moon, Lewis Research Center took special pride in the fulfillment of Kennedy's promise to land Americans on the Moon within a decade. Armstrong had started his career in Cleveland as a NACA test pilot. Lewis engineers had helped solve some of the problems of Saturn V's huge F-1 engines, and tests in the Altitude Wind Tunnel's vacuum chamber had helped to pave the way for the safe separation of the Apollo capsule from its boosters.58 The Ranger spacecraft, the Lunar Orbiter, and Surveyor - launched by Agena and Centaur - had also contributed to the success of Apollo, but the flight of Apollo across the Sun had cast a long shadow. The nation's inability to deal with its social problems seemed to make a mockery of ambitious plans for space exploration.

For Silverstein 1969 marked not only the achievement of a manned lunar landing, but also 40 years of government service. Silverstein's technical leadership had shaped NASA's early years, but Headquarters was now in hands more attuned to the political winds than to charting a post-Apollo course to the planets. In his dealings with Headquarters, Silverstein became like a cactus with thorns. 59 True to the NACA ideal of cultivating expertise from within the laboratory, he fought NASA's increasing commitment to outside contracting at the expense of research. By 1969 all three former NACA laboratories - Lewis, Langley, and Ames - were outside NASA's mainstream. With the exception of the laboratory's management of unmanned launch vehicles, which included Titan and Atlas in addition to Centaur, Lewis received its funding through the Office of Aeronautics and Space Technology (OAST, formerly OART). OAST commanded a mere 5 percent of NASA's budget. With manned missions dominating the agency, the camel had forced its research occupants nearly out of the tent. The research centers were looked down on by other parts of NASA as "hobby shops."60 Asked to chair a study for the expansion of Hopkins International Airport into Lake Erie, Silverstein chose to retire. It was the end of an era.

 


Abe Silverstein, Director of Lewis Laboratory from 1961 to 1969.

Abe Silverstein, Director of Lewis Laboratory from 1961 to 1969.

 


Notes

1. Cleveland Plain Dealer, 13 December 1961.

2. Abe Silverstein and George R Kinghorn, "Improved Baffle Designs for Air-Cooled Engine Cylinders," August 1943, issued as NACA WR L-767. Abe Silverstein, "Internal Aerodynamics of Ramjets," material presented at the NACA Conference on Supersonic Aerodynamics at AAL on 4 June 1946, NACA 1946/1 1475-9.

3. "Big Joe", Lewis' Part in the Project Mercury Story," Orbit, 22 May 1959. Lloyd S. Swenson, Jr., James M. Grimwood, and Charles C. Alexander, This New Ocean: A History of Project Mercury, NASA SP-4201 (Washington, D.C.: U.S. Government Printing Office, 1966), p. 201 and 244-245.

4. T. Keith Glennan, "The First Years of the National Aeronautics and Space Administration," 1964, unpublished diary, Eisenhower Library, Abilene, Kan., p. 84,

5. New York Times, 12 March 1960. Silverstein biography file, NASA History Office, Washington, D.C.

6. Interview with Simon Ostrach, 29 September 1987.

7. See, for example, the description of the NACA by Homer Newell, Beyond the Atmosphere, NASA SP-4211 (Washington, D.C.: U.S. Government Printing Office, 1980), p. 90-91.

8. T Keith Glennan, "The First Years," vol. 1, p. 7

9. Ibid.

10. E. J. Manganiello, "The Changing Trends in Our Activities," address to staff, 4 November 1960. RG 255 73A32, 116/1-52, Federal Archives and Record Center, Chicago, IL.

11. Ibid.

12. Ibid.

13. Ibid.

14. Quoted in Arthur L. Levine, "United States Aeronautical Research Policy 1915-1958: A Study of the Major Policy Decisions of the National Advisory Committee for Aeronautics," Ph.D. Dissertation, Columbia University, 1963, p. 90.

15. United States Civilian Space Programs, 1958-1978, Report Prepared for the Subcommittee on Space Science and Applications, Serial D, Volume I, January 1981, p, 72.

16. Interview with Simon Ostrach, 29 September 1987.

17. Interview with Bruce Lundin, 28 May 1987.

18. Cleveland Plain Dealer, 13 December 1961.

19. See Daniel S. Greenberg, The Politics of Pure Science (New York: The New American Library, 1967), p. 258 259.

20. Courtney G. Brooks, James M. Grimwood, and Lloyd S. Swenson, Jr., Chariots for Apollo: A History of Manned Lunar Spacecraft, NASA SP-4205 (Washington, D.C.: US. Government Printing Office, 1975), p. 46-47. See discussion in Roger E. Bilstein, Stages to Saturn, NASA SP-4206 (Washington, D.C.: U.S. Government Printing Office, 1980), p. 60-68.

21. Brooks, Grimwood, and Swenson, Chariots for Apollo, p. 34.

22 "Survey of Various Vehicle Systems for the Manned Lunar Landing Mission," Bruce Lundin, Chairman, 10 June 1961. NASA History Office, Washington, D.C.

23. Interview with Abe Silverstein by V. Dawson, 6 June 1987.

24. Interview with Bruce Lundin, 28 May 1987. On a lunar excursion vehicle, see Carl R Schueller to Associate Director, "Visit to NASA Headquarters on November 23, 1959 by Mssrs. E. W. Conrad and C. F. Schueller!' See also Richard J. Weber to Chief, Propulsion Aerodynamics Division, "Lunar Mission Studies by Analysis Branch, Propulsion Systems Division," 9 June 1960, personal files of R. J. Weber.

25. See discussion by Arthur M. Squires, The Tender Ship. Governmental Management of Technological Change (Boston: Birkhauser, 1986), p. 94-102. The classic study of nuclear propulsion in all its facets, both scientific and technical, was published in the open press by two British scientists, L. R. Shepherd and A. V. Cleaver, in the journal of the British Interplanetary Society (Sept.-Nov. 1948 and Jan.-March 1949). Thereafter, technical interest in nuclear-powered rockets languished because of the enormous technical problems they seemed to pose,

26. NACA Press Release, 26 September 1956, NASA Lewis Records.

27. See "Remarks of Melvin Price," Congressional Record Appendix, 26 August 1960.

28. See R. W. Bussard and R. D. DeLauer, Fundamentals of Nuclear Flight (New York: McGraw Hill, 1965), p. 1-4.

29. For background see R. E. Schreiber, "Kiwi Tests Pave Way to Rover." Nucleonics, vol. 19, No. 4, April 1961, 11.77-79.

30. Glennan, "The First Years," p. 165.

31. See Harold B. Finger, "Space Nuclear Propulsion Mid-Decade," Astronautics & Aeronautics, January 1965. The Plum Brook reactor went critical in 1961 and reached full capacity in 1963. The reactor was shut down on 30 June 1973. By the 1970s Plum Brook had a staff of 635 civil servants and 132 support service contractors.

32. Glennan, "The First Years," p. 71.

33. For background on NERVA, see W. H. Esselman, "The NERVA Nuclear Rocket Reactor Program," Westinghouse Engineer, May 1965, vol, 25, p. 66 75. See also William R. Corless, Nuclear Propulsion for Space (U.S. Atomic Energy Commission, 1967). See also James Arthur Dewar, "Project Rover: A Study of the Nuclear Rocket Development Program, 1953-1963," Ph.D. Dissertation, Kansas State University, Manhattan, Kansas, 1974, to be published by Smithsonian Institution Press.

34. W. E. Moeckel, "Status of Electric Propulsion Systems for Space Missions," Cryogenic Engineering Conference, Boulder, Col., 23-25 August 1960, RG 255 73A32 116/1-52, Federal Archives and Record Center, Chicago, IL.

35. R. E. English, H. O. Slone, D. T. Bernatowicz, E. H. Davison, and S. Lieblein, "A 20,000 Kilowatt Nuclear Turboelectric Power Supply for Manned Space Vehicles," NASA RM 2-20-59E, March 1959.

36. W, E. Moeckel, "Lewis Research Center Electric Propulsion Program," NASA Headquarters Presentation, 5 June 1961, RG 255 73A32 116/1-52, Federal Archives and Record Center, Chicago, IL.

37. H, R. Kaufman, "Origin of Electron-Bombardment Ion Thruster," J Spacecraft, 18:289-292.

38. "SERT II-A First for Lewis," Lewis News, 16 January 1970.

39. John Sloop, "Comments by John L. Sloop on Manuscript History of the Lewis Research Center by Virginia Dawson," 5 September 1989. NASA History Office, Washington, D.C.

40. Brian Sparks to Homer Newell, 21 September 1962, Atlas-Centaur Launch Vehicle file, NASA History Office.

41. "Review of the Atlas-Centaur Launch Vehicle Development Program," Report to the Committee on Science and Astronautics, House of Representatives, March 1963, Atlas-Centaur Launch Vehicle file, NASA History Office.

42. See also files of Office of the Deputy Director 0100, Centaur Program, NASA Lewis Archives, Box 9.

43. Interview with Harlan Simon, 20 March 1985.

44. Interview with Bruce Lundin, 28 May 1987.

45. Interview with Harlan Simon, 20 March 1985.

46. Ibid.

47. Ibid.

48. Photo from personal files of Abe Silverstein.

49. Arthur M. Squires discusses "maestros of technology" in The Tender Ship, p. 13-46. See also Abe Silverstein and Eldon W. Hall, "Liquid Hydrogen as a jet Fuel for High-Altitude Aircraft," NACA RM E55 C28a, 15 April 1955.

50. For greater detail on the Agena project, its management and significance, see Seymour Himmel, "Commentary" [on manuscript by V. Dawson], 28 August 1989.

51. "A Proposal to the National Aeronautics and Space Administration for Support of Materials Research and Departmental Facilities at Case Institute of Technology," 8 April 1960. Case Central File 19 DC: 27: Asso. and Orgs. Gov. Gps. & Ind. Orgs., Case Western Reserve Archives.

52. T Keith Glennan to Dr. John Hrones, "NASA Support for Research Proposals," 8 October 1962, Case Central File 19 DC 27:2, Case Western Reserve Archives. Glennan also protected the integrity of the memo. It was not to be found in the Case archives.

53. R. H. Thomas to Dr. T. Keith Glennan, "Visit by Dr. T. L. K. Small," 1 October 1964; and R. H. Thomas to J. A. Hrones, "NASA Visitor-Sid Roth:' Case Central File 19 DC 27, Case Western Reserve Archives.

54. "Conference on Proposed Electronics Center," 19 November 1963, Case Central File 19 DC 27:2, Case Western Reserve Archives. Also Abe Silverstein to Francis B. Smith, Area Survey Committee, 6 December 1963, NASA Lewis Records, 010019.

55. See Selected Technology for the Electric Power Industry, a conference held at Lewis Research Center, Cleveland, Ohio, 11-12 September 1968, NASA SP-5057, 1968.

56. "Proposal to the National Aeronautics and Space Administration for Case Laboratory for Space Engineering Research," 27 January 1965, Case Central File 19 DC 27: 1, Case Western Reserve Archives.

57. NASA Historical Data Book, NASA SP-4102 (Washington, D.C.: U.S. Government Printing Office, 1988), vol. 1, p. 166 and 168.

58. "Apollo Mans Moon," Lewis News, 18 July 1969.

59. Interview with Robert E. English by V. Dawson, 11 July 1986.

60. Ibid.


previousindexnext