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

 

CHAPTER EIGHT

SEIZING THE SPACE INITIATIVE

 

[149] There appears to be a strong parallel between the NACA's transition from aircraft piston engines to jet propulsion and the transition from air-breathing engines to rocket propulsion.1 In both cases the NACA failed to anticipate the revolutionary change in the nation's propulsion requirements. However, the historical context in which these transitions occurred was different. The American propulsion community in the late 1930s ignored jet propulsion because they were caught up in the peacetime development of commercial engines. Just the opposite was true in the early 1950s. The same rocket intended to be lobbed at the Soviet Union could propel a satellite into orbit around Earth, but national security dictated the nation's propulsion priorities. Until 1955, when planning for the International Geophysical Year began, space was dismissed as science fiction. The military focused its energies during this period on missile development, and the NACA remained, willingly or unwillingly, on the sidelines. The launch of Sputnik by the Soviet Union in 1957 caught the nation, including the NACA, off guard.

Why the NACA stayed out of the mainstream of large rocket development may be as much a political question as a technical one. Hugh Dryden never went to Congress without a carefully conceived, down-to-earth program. He was all the more conservative when it came to asking Congress for funding outside of the NACA's specific aeronautics mandate. In the political climate of the early 1950s it was increasingly difficult to get the budget-minded Congress to support science and technology appropriations. The military requested and obtained funding for research folded into large budgets justified on national security grounds. This generous funding enabled the Air Force to build new research facilities at Wright Field and an array of the latest in wind tunnels at Arnold Engineering Center in Tullahoma, Tenn.

The Army, Navy, and Air Force found themselves with ample financial resources for missile development. Given the constraints on NACA funding prior to Sputnik, from the point of view of one NACA engineer the NACA "would have stood as much chance of injecting itself into space activities in any real way as an icicle in a rocket combustion chamber."2 Yet the question remains. Could an earlier and more sustained attention to rocket research have positioned the NACA on the propulsion frontier before space missions captured the nation's imagination and forced Congress to open its pocketbook?

Between 1945 and the early 1950s research at Lewis Laboratory focused on improving the turbojet, a bulky, roaring, fuel-thirsty engine, to a quiet, dependable, commercially viable propulsion system. However, this program in support of existing technology did not preclude more advanced work. Beyond the duty to respond to specific requests from the military and industry [150] there was sufficient autonomy and flexibility within the program of research for projects to grow organically from within the laboratory. It was possible to allow work on less conventional projects in a few well-chosen areas. If a particular project won the attention of Abe Silverstein, it received support. Silverstein's "pet projects" may have had only a small percentage of the total budget of the laboratory, but he assigned some of his most talented staff to these projects and watched over their progress with an attentiveness that the more routine projects did not receive. The problem for the members of the rocket section wag how to win the attention of Lewis's upper management. At first, they were frustrated by the lack of support for their work.

Since 1945, when the Cleveland laboratory established its rocket section, its small staff had continued to study rocket fuels despite the general attitude of headquarters and Lewis upper management against a strong NACA role in rocket development. The group's work was limited to the study of propellants, particularly high-energy fuels, thrust chambers of rocket engines, combustion, and cooling. By 1948, the rocket group had produced calculations on the performance of a number of propellants. Among liquid propellants worthy of experimental study, liquid hydrogen appeared to hold greatest promise because of its high specific impulse.3 The rocket group became convinced that the potential of liquid hydrogen to produce greater thrust than conventional fuels could offset the disadvantages of its low density. However, the use of liquid hydrogen as a rocket fuel presented enormous technical problems. Liquid hydrogen is a cryogenic fuel, dangerous to handle and difficult to store. To verify predictions, laboratory facilities were needed. In addition, it was difficult to obtain an adequate supply of the fuel for experiments. The study of liquid hydrogen had to be passed over in favor of fuels like hydrazone and diborane to be used with oxidizers like hydrogen peroxide, chlorine trifluoride, liquid oxygen, nitrogen tetroxide, and liquid fluorine.4

When Abe Silverstein became Chief of Research, he allowed the rocket group more responsibility and visibility. In 1949 the rocket section became a branch within the Fuels and Combustion Research Division, headed by Ted Olson. John Sloop, head of the new branch, recalled, "it was moved up one level in the organizational hierarchy, named for what it was, and given more personnel."5 However, in the early 1950s, rocket research remained a small fraction of the work of Olson's division, which focused on the combustion problems and fuels of turbojets, and ramjets. Silverstein's interest was limited to encouraging the group to establish the criteria for propellant selection. Silverstein also became mildly interested in liquid hydrogen, perhaps in connection with nuclear rocket propulsion. A 1947 secret report by physicists at the Applied Physics Laboratory at The Johns Hopkins University had proposed hydrogen as the preferred propellant for a nuclear rocket.6 Interest in hydrogen-fueled rockets thus dovetailed with the increasing emphasis on nuclear aircraft and rocket propulsion on the part of the Air Force. Liquid hydrogen also had possibilities as a fuel for high-altitude aircraft such as the U-2. Silverstein, however, did not lend any muscle to support requests for additional facilities. Without facilities, the experimental side of the research suffered.

Members of the rocket branch realized that, to change the attitude of Lewis management from polite tolerance to firm commitment, they had to do more than produce their quota of research papers. They became advocates for increased attention to rocket research to audiences both within and outside the laboratory. Although they were careful not to push the space applications too hard, other Lewis staff could not resist kidding them for becoming "Buck Rogers types." Sloop, for example, gave two lectures in 1949 as part of the Case extension course (ME 221) offered at Harding junior High School. The first lecture tackled the general problems of rocket [151] engines and performance, including regenerative cooling; the second, solid propellant rockets. Sloop's references reveal his command of the groundwork established by rocket pioneers Robert Goddard, Maurice Zucrow, Willy Ley, and Frank Malina. Sloop indicated that, beyond missile applications, step rockets could be used to escape Earth's gravity. He went so far as to suggest the establishment of a space station.7 At local, regional, and national meetings of the Society of Automotive Engineers ISAE), Kiwanis Clubs, and American Legion Posts, Sloop and his staff hammered away at the same themes: the potential of rocket-powered missiles both as military weapons and as research tools to gather "basic scientific data about the upper atmosphere and interstellar space." They advocated "a vigorous, large-scale research and development program."8

In 1951, possibly in response to intelligence reports of Russian advances in rocketry, the NACA authorized a formal Subcommittee on Rocket Engines within the Power Plants Committee. That same year Lewis Laboratory received its first formal appropriation for rocket research. However, the number of personnel assigned to the rocket branch was still small, less than 3 percent. The experimental facilities consisted of four rocket test cells constructed during World War II and four larger test cells put up by an ad hoc laboratory construction group, paid for from laboratory operating funds. The laboratory still lacked adequate facilities for production, storage, and testing of liquid hydrogen rockets.9

Members of the rocket branch staff paid close attention to Army missile policy, which in 1952 was beginning to swing away from air-breathing propulsion systems. The rocket group began to plan a large rocket engine test complex for a remote location in the West, later scaled down to a single facility appropriate for a site at Lewis. In 1953 the laboratory acquired a hydrogen liquefier, but it had to wait until 1957 for the new high-energy rocket propellant test facility to be ready for operation.

In its quest for a large-scale rocket facility, the rocket group seems to have received greater encouragement from the new NACA rocket engine subcommittee than it did from either Dryden or Lewis management. John Sloop lamented:

Ironically, it was the special subcommittee on rockets set up in 1951 and managed by NACA headquarters that was the most influential in prodding both headquarters and laboratory management into doing more research on rocket engines. The recommendations of the subcommittee, and particularly those of Chairman Maurice Zucrow, were crucial in getting the $2.5 million rocket propulsion laboratory for high-energy propellants, with construction beginning in 1953.10

It is possible that the tarnished reputation of the NACA in the early 1950s made Dryden reluctant to push for an expanded role in the national program of rocket research. In 1951 the....

 


Rocket Test Cell, 1952.

Rocket Test Cell, 1952.


[
152]

Diagram of a Liquid Propellant Rocket, 1953.

Diagram of a Liquid Propellant Rocket, 1953.

 

....NACA was jolted by the arrest of William Perl. Perl, a brilliant theoretician, had worked at Langley before his transfer to Lewis. The previous year he had completed his Ph.D. under Theodor von Karman. He was accused of perjury in connection with the sensational trial of Julius and Ethel Rosenberg. Perl had denied to a Grand jury that he knew the Rosenbergs when, in fact, during his year as a graduate student at Columbia University, he had sublet an apartment rented to one of the alleged conspirators. The FBI suspected that Perl had communicated sensitive aerodynamic information, including national plans to develop a nuclear powered airplane, to the Soviet Union through the Rosenbergs. These allegations were never proved. Perl was convicted of perjury and sentenced to five years in prison. Whether there is a connection between the Perl case and the NACA's funding difficulties after the Korean War is pure speculation. At the height of the nightmare of McCarthyism, it does not seem surprising that the NACA might have been considered a security risk."11

The NACA's woes were compounded in January 1952 when Albert Thomas, Chairman of the House Appropriations Subcommittee, demanded an investigation of the NACA by the General Accounting Office (GAO), the first such scrutiny the NACA had ever experienced. Thomas thought that the agency had too many civil service employees and possible duplication of research effort among the three laboratories. In September the GAO sent its auditors, followed by a visit from Congressman Thomas himself. One can imagine the trepidation with which laboratory management prepared for his visit. After an extensive tour, including stock rooms and new facilities, Mr. Thomas, it seems, went away favorably impressed.

[153] At the Materials and Stresses Building he heard about the division's work in high-temperature materials and its progress in developing alloys and ceramal turbine blades. At the Propulsion Systems Laboratory still under construction, Sharp stressed the utility of NACA testing to support turbojet development. He reported to Dryden:

 

He was impressed by the possibility of getting information on these engines before they went into production, and I think he realized for the first time the potentialities of this work since he commented that it was obvious that large savings could be made if the engines were proven before they were put into production. Over lunch Silverstein turned the discussion to the NACA's future research. He stressed "the need for more intensive research and development on the rocket engine in this country in a connection with the long-range missile." He let Thomas know that a request for new facilities for rocket development was part of the 1954 budget request.12

 

FROM MACH 4 TO INFINITY

 

In July 1955 President Dwight Eisenhower announced that the United States would participate in the International Geophysical Year (to begin in 1957). The United States planned to launch a satellite. The idea of space exploration now caught the imaginations of men higher up in the laboratory hierarchy - John Evvard, George Low, and Wolfgang Moeckel. Aware of the work in hypersonics at Langley and Ames, they thought that both space flight and propulsion problems related to aerodynamic heating should be explored.

Preparations for the International Geophysical Year sparked a new awareness and commitment within the scientific community to the peaceful development of space sciences. Possibly because of its scientific imprimatur and the fact that it was to be managed by a team of Americans, the relatively weak joint proposal by the Vanguard group at the Naval Research Laboratory and the National Academy of Sciences beat a proposal by the more experienced von Braun team.

Lewis Laboratory's preparations were more modest. To prepare the staff for the problems of flight beyond Earth's atmosphere, they began to consider what courses to offer the following year in the non-credit system that flourished under the direction of John Evvard. Because the course system served as a bellwether for new research initiatives and future reorganizations of the laboratory, Evvard carefully laid the groundwork through discussions with possible instructors. With considerable trepidation Evvard went to see Silverstein. In describing the proposed course, Evvard predicted that it could "change the thinking of the people who take the course," something that might not be desirable because of the dissatisfaction it could cause with more routine work.13 They planned to call the course "From Mach 4 to Infinity." Without hesitation Silverstein gave Evvard his support. "From Mach 4 to Infinity" created a strong impression on those who took it in 1956. The second year it was offered, the course was enormously popular.

A respected group composed mainly of the laboratory's theoreticians - Melvin Gerstein, George Low, Roger Luidens, Stephen Maslen, and Harold Mirels - developed the part of the course that dealt with the concepts of hypersonic flight. Some basic research in heat transfer by Clarence B. Cohen, Eli Reshotko, and Ernst Eckert provided a useful starting point.14 Wolfgang Moeckel began to explore the idea of possible flights to the planets using unconventional propulsion systems. As part of his lecture "Earth Satellites and Interplanetary Travel," Moeckel reviewed the literature on rocket propulsion. He recalled that he had first come across a description of [154] electric rocket propulsion in about 1949 in Hermann Oberth's 1929 book, Wege zur Raumschiffahrt. He was particularly intrigued by a chapter devoted to the electric rocket that suggested that, for space travel, rocket thrust could be produced by the flow of electrically charged particles.15

The chemical rocket was limited by the enormous amounts of propellant required to be carried into space for flight to distant planets. Although chemical rockets produced far greater thrust, they had to carry large quantities of fuel into space. Electric rockets got their power from a generator in space providing very small amounts of thrust over very long periods of time. This suggested to Oberth, as it had to Robert Goddard in 1906, that electric rockets might be useful for long-distance travel between planets. However, neither Oberth nor Goddard tackled the practical questions involved in the development of a power source. The electric rocket was consigned to the category of intriguing but impractical technical ideas.16

Moeckel's colleagues received his first lecture with considerable interest. Expanding these lectures later the same year, he came across two papers by Ernst Stuhlinger, who had worked on the V-2 at Peenemunde under Werner von Braun.17 Moeckel may have heard about Stuhlinger's work at the semi annual meeting of the American Rocket Society held in Cleveland in 1956 where Krafft Ehricke, also of the Peenemunde group, lectured on "The Solar-Powered Space Ship."18

Stuhlinger had become intrigued with the possibilities of electric propulsion as early as 1947 at the Army Camp at Fort Bliss when von Braun asked him to restudy Oberth's rocket work. Stuhlinger's first paper, published in 1954, tackled the question that Oberth had left unanswered: how to generate the necessary electric power in space. He suggested a solar turbogenerator that consisted of a system of mirrors and boilers. In his second paper, published in 1955, Stuhlinger took a more radical approach. He suggested a nuclear fission reactor as the power source. He reasoned that "a vehicle designed for a Mars mission would be lighter and somewhat faster if powered by a nuclear reactor than if it were powered by solar energy."19

Stuhlinger's papers impressed Moeckel because they were "much more concrete, comprehensive and realistic than previous work." Once Stuhlinger had suggested that electric propulsion was within the realm of technical feasibility, Moeckel set enthusiastically to work. "I immediately began a study of low-thrust trajectories, and the capabilities of such low thrust systems for interplanetary travel."20 If headquarters had been apprised of Moeckel's work, it might have been discouraged as "space-cadet" enthusiasm. Indeed, it has yet to find application in space exploration. However, like the analytic work done on liquid hydrogen before large experimental facilities were authorized, or the theoretical work on hypersonics, there was no need to inform Washington at this point. Moeckel received the strong support of his division chief, John Evvard, then head of the Supersonic Propulsion Division, who gave him the necessary freedom to think through his ideas.

Meanwhile, Walter Olson began to make a concerted effort to influence Dryden to take a more aggressive approach to NACA rocket research. In "A Suggested Policy and Course of Action for NACA with Regard to Rocket Engine Propulsion (1955)," Olson advocated "advancing the engine art" by experimental work on rocket engines of 250,000 to 1,000,000 pounds thrust. Although missile applications are clearly intended in the document's careful wording, Olson, almost apologetically, urged the NACA to also consider satellite flight "which stiff seems visionary to many minds now:' Nuclear-powered aircraft, he suggested, might serve as launching pads for flight beyond the atmosphere. Rockets could also be used for auxiliary thrust to increase the speed of lumbering nuclear-powered aircraft.21

[155] Although Dryden still demurred, Silverstein began to take a more active interest in the work of the rocket group on liquid hydrogen. The stimulus came unexpectedly from aeronautics. The Air Force, interested in liquid hydrogen to fuel turbojet engines for high-altitude reconnaissance aircraft, supported an experimental test program at Lewis Laboratory called Project Bee. Silverstein put Paul Ordin in charge of the project and made Donald Mulholland his assistant. This was exactly the kind of hands-on engineering project that Silverstein found most challenging. At last the work of the rocket branch in high-energy propellants had gained his full attention. He set up a room in the basement of the Administration Building directly below his office for the project so that he could personally direct it. The project staff modified a B-57B bomber equipped with two Curtiss Wright J-65 engines so that one engine could burn either jet fuel or hydrogen. After extensive ground testing of the hydrogen system, it was flight tested over Lake Erie. When the pilot switched the modified engine to hydrogen fuel, it worked perfectly. The program resolved the questions of insulation, structure, and pumping of liquid hydrogen in flight.

Project Bee validated the work of the rocket team, as well as proving the feasibility of liquid hydrogen as a fuel for aircraft. Through Project Bee, Lewis researchers received important practical experience in the handling and storage of liquid hydrogen that would later prove crucial when the laboratory took charge of the development of the Centaur rocket. Abe Silverstein, with Eldon Hall, wrote the final report: "Liquid Hydrogen as a jet Fuel for High-Altitude Aircraft."22 The laboratory set up an engine design group led by William D. Ritter to study liquid hydrogen turbojet engines for supersonic flight.23

In July 1956 Dryden at last began to reconsider his opposition to increased NACA participation in rocket research. He took the belated initiative to canvas the opinion of the NACA rocket subcommittee to determine the NACA's long-term research agenda. This conservative approach was typical of headquarters when entering a new area of research. Dryden's cautious modus operandi was to "solicit opinions and build a broad base of national support so that it would appear the agency was practically pushed into the new work."24 For five years, Dryden had ignored both the rocket subcommittee's recommendations and pressure from Lewis Laboratory to allow more effort in rocket research. On the behalf of frustrated members of the NACA rocket subcommittee, Richard Canright (employed by Douglas Aircraft after leaving the jet Propulsion Laboratory) wrote in July 1957:

 

We have constantly spurred the NACA on to tests on a larger scale. We have urged them to become familiar with complete engines rather than work only on component R&D. We have tried to emphasize the importance of rocket technology to this country's defense effort and urged that the NACA devote a greater portion of its personnel and funding to this important. field.25

 

On the eve of Sputnik, Dryden still hesitated to take up the recommendations of the rocket subcommittee, roll up his sleeves, and challenge a tight-fisted Congress for a larger NACA role in rocket development. Nevertheless, Lewis management was now solidly behind the work of the rocket group.

 

TRANSITION TO SPACE

 

Late in 1956 Abe Silverstein was ready to leave the problems of the turbojet engine to industry. Two new facilities under construction - the rocket engine test facility and the nuclear reactor at Plum Brook - required a rethinking of the laboratory's research program. Silverstein began [156] to consider a major reorganization of the magnitude of the one that had accompanied the switch from the aircraft reciprocating engine to jet propulsion in 1945.

Silverstein selected some of the Compressor and Turbine Division's most versatile engineers and scientists to attend a "nuclear school." Eight additional engineers were plucked from other divisions of the laboratory and invited to drop their current projects and prepare for leadership in the anticipated reorganization by learning the principles of nuclear physics. They were taught by the small core of researchers at the laboratory who had been engaged in nuclear work for almost a decade and professors invited from Case Institute of Technology. In all, 24 engineers had no duties other than to attend the school for six months. The rationale for the nuclear school appears to have been to single out individuals with leadership qualities and teach them the fundamentals of new areas Silverstein believed would become the major focus of research in the future. Presumably, these individuals would be the division and branch chiefs after the reorganization. In January 1957 six engineers chose not to continue in the nuclear field. The remaining 18 were divided into three groups. Harold Finger headed a group to work on a nuclear rocket. In 1960 he would be selected by Silverstein to become project director of the joint Atomic Energy Commission-NASA nuclear projects at headquarters. Eldon Hall formed a group to study aircraft nuclear propulsion. He would also go to Washington with Silverstein after NASA was organized. Robert English focused the energies of his group on a design study for a nuclear space-power system. With the imagined goal of sending eight people to Mars, they proposed a nuclear reactor with the potential to act as the power source for an electric rocket under consideration by Moeckel's group.26 Robert Graham, head of the Rocket Fluid Dynamics Section, took charge of planning the laboratory's first seminar on rocket propulsion to train additional staff in rocket technology.27

In March 1957 Silverstein's first act in shifting the laboratory's research priorities was to establish a research planning council. He abolished the Compressor and Turbine Division, long one of the premier divisions of the laboratory. Many of the aerodynamicists who had worked on the axial compressor were assigned to a new Fluid Systems Division to study the mechanics of flow within rocket systems. At the same time Silverstein created a Nuclear Reactor Division to direct new research connected with the reactor then under construction at Plum Brook.

 

 


A B-57 aircraft had one engine fueled with liquid hydrogen.

A B-57 aircraft had one engine fueled with liquid hydrogen.

 

[157] Meanwhile, the streams leading to space research were converging. In spring 1957, the staff began to plan for a fall conference on the evaluation of propulsion systems. Proponents of space-related research urged Silverstein to allow them to include sessions on space flight and space propulsion systems, as well as more conventional subjects, such as turbojets and ramjets. Never one to avoid considering unconventional technical concepts if there were a possibility they might ultimately prove feasible, Silverstein agreed to give them one afternoon session. With the imagination of many staff now sparked, research groups worked up sessions on liquid hydrogen, nuclear rockets, and space power systems. As Moeckel recalled, "The studies for this conference formed the foundation for the rapid expansion of our research in the fields of electric propulsion, power generation, and nuclear rockets." The paper submitted by Moeckel's group on "Satellite and Space Propulsion Systems" contained a distillation of the advanced thinking of the laboratory-flights to the Moon and Mars, with the focus on electric propulsion systems.28

As preparations for the propulsion conference moved into high gear in September 1957, Silverstein called a meeting of the Research Planning Council, whose members included Eugene Manganiello, John Evvard, Bruce Lundin, Walter Olson, Irving Pinkel, and Newell Sanders. He expressed the consensus of the group that "existing problems on the air-breathing turbojet engine were not of sufficient laboratory interest for the continuation of a large-scale program."29 His proposal that rocket research be expanded, while proportionately reducing turbojet engine research, effectively marked Lewis Laboratory's transition to space research. However, as the November conference approached, Silverstein became increasingly nervous about possible criticism by Headquarters of the inclusion of a discussion on space propulsion. Would the laboratory be perceived as an amateur group of space-cadets? Not willing to eliminate the session entirely, he cut the panels from a full afternoon to several hours.

 

RESPONSE TO SPUTNIK

 

The tension between the foot-dragging (if headquarters and the gathering momentum of the Cleveland laboratory toward space-related research mounted in September when representatives from headquarters visited the laboratory to evaluate its rehearsal for the finely orchestrated NACA Triennial Inspection. The theme of the inspection was to be a celebration of the tenth anniversary of the X-1, the first aircraft to fly faster than the speed of sound.30 Unlike the anticipated more specialized propulsion conference directed toward a technical audience, NACA inspections were intended to convey its work in layman's terms to a group made up largely of politicians and industry executives. At one of the "stops," the rocket group proudly showed John Victory the new rocket laboratory from a small platform next to the huge new scrubber, part of the silencing and exhaust gas disposal system. During the prepared talk, Victory bristled when he heard references to space. Always on the lookout for anything that might offend potential Congressional sponsors, he ordered all references to space deleted from the presentations. As one member of the rocket group explained, "The climate in Washington in the fall of 1957 was very negative towards space." It was acceptable to mention the "slow-paced" Vanguard satellite managed by the Navy under the aegis of the International Geophysical Year, but "anything beyond it was considered 'space cadet' enthusiasm."31 Nevertheless, Addison Rothrock, who was also in the headquarters contingent, was able to convince Victory to allow one of the first electric propulsion experiments - a rail accelerator - to be included in the presentations.

The intended celebration of ten years of supersonic flight caught the NACA looking backwards. On October 4 the question of whether to discuss space-related work was moot.

[158] Between the rehearsal and the actual inspection, the Soviet Union launched Sputnik, the world's first artificial satellite. Suddenly Chuck Yeager's dramatic breaking of the sound barrier seemed like ancient history. Sputnik's flight beyond the atmosphere marked the dawn of the space age, a new era of discovery. However, to Americans, Sputnik also seemed ominous. Politically, the new satellite was a symbol of the rising technical competence of an enemy. In the context of the Cold War, Sputnik represented the Russian triumph in the first round of what immediately began to be perceived as a space race. Victory's earlier resistance to references to space put the conservative attitudes of headquarters in sharp relief. When the inspection began several days later, Lewis engineers proudly unveiled their work on chemical rockets and more visionary space propulsion systems.

The presentations on high-energy rocket propellants were the highlight of the inspection. Participants could admire a rocket engine capable of 20,000 pounds of thrust ready for experiments with liquid fuels when they visited the new rocket engine test facility.32 Sputnik at last riveted the attention of the laboratory on work of the small contingent of "Buck Rogers types." They could describe with impunity an idea for a winged satellite, similar to the shuttle, lofted beyond the atmosphere by a multi-stage rocket booster. It was a day that vindicated their long commitment to rocket engines and fuels. They found themselves besieged by their colleagues in air-breathing propulsion for briefings on rocket fundamentals.

The most detailed consideration of Lewis Laboratory's work in space propulsion was reserved for the classified NACA-Industry Conference, held the following month. Once again, the rocket researchers held center stage. The presentation by John Sloop, A. S. Boksenbom, S. Gordon, R. W. Graham, R M. Ordin, and A. O. Tischler discussed propulsion requirements for specific missions, including surface-to-surface missiles, Earth satellites, and Moon missions. They considered both circumnavigation of the Moon and an ambitious Moon landing, using an orbiting Earth satellite as a base, probably the first detailed discussion of a Moon landing in NACA literature. Frank E. Rom, Eldon W, Sams, and Robert E. Hyland's paper, "Nuclear Rockets," and the paper "Satellite and Space Propulsion Systems," by W C. Moeckel, L. V. Baldwin, R. E. English, B. Lubarsky, and S. H. Maslen, were equally visionary.33 John Sloop and his colleagues had at last found a receptive audience for their stubborn and lonely advocacy.

 

THE SPACE DEBATE

 

During the national soul searching that followed the Russian triumph in space, Lewis staff began to consider their role in the charged political and technical environment. Some engineers looked forward to abandoning air-breathing engines to tackle the problems of engines in zero gravity. For others, to exchange the familiar roar of the wind tunnel for the silence of the vacuum chamber seemed a travesty. Bruce Lundin recalled that half the laboratory was afraid of getting "sucked" into space and the other half was afraid of being left out. "And there was some concern among people, whose views I did not personally share at the time, that if we got into space we'd be into an operating mission agency and the good things that they were doing in research would be lost.34

Discussion of the NACA's role in space dragged on through meetings of the Lewis Research Planning Council and in informal debates in the cafeteria. This prompted the scholarly Walter Olson to draft a document in support of space flight research. He pointed out that space flight had important scientific and military applications. Space missions would benefit meteorology and astronomy by yielding new data on radiation, meteorites, gas composition, electromagnetic [159] phenomena, and cosmic dust in space. Reconnaissance was an obvious military application. Olson also stressed the propaganda value of space exploration. It would not only be a means to demonstrate the technical superiority of the United States - the importance of having the "technological capability of massive retaliation" but it would also keep "both friend and potential foe convinced that such is the case."35 Olson listed 15 specific problem areas in which he thought the NACA could make immediate contributions to a new space initiative. Most would be collaborative efforts with other NACA centers or existing government agencies. They included space propulsion systems such as chemical rockets, nuclear rockets, and the study of ion, plasma, and photon jets; auxiliary power systems; and materials for space vehicles and exploration equipment. He advocated a new NACA laboratory to launch and manage a manned, orbiting, space platform. In his view this manned platform should be the main focus of NACA activities in space.

This plan was not bold enough for Bruce Lundin, although many of Olson's ideas were incorporated into Lundin's plan. Lundin thought that the NACA should aim at nothing less than leadership of an entirely new national space agency. At home on a quiet Sunday afternoon in early December, he produced a memo for Abe Silverstein that he called "Some Remarks on a Future Policy and Course of Action for the NACA." He argued against the collaborative approach, which he thought "weak and ineffective."36 He advocated a "bold, imaginative, aggressive, and visionary" program. He warned that if the NACA focused on a specific project, like a manned space platform or placing a dye marker on the Moon, this might "dangerously limit our goals, restrict the range of our thinking, and give us nothing to grow on." Aeronautical research, not directly controlled by the military and directed to national goals, had been the traditional role of the NACA. In tones resonating with Cold War rhetoric, Lundin declared that space research was a matter of national survival:

 


The Rocket Engine Test Laboratory completed in 1957.

The Rocket Engine Test Laboratory completed in 1957.

 

 

[160] In our technological age, it will be the country that advances in science that will have the greatest impact on the emotions and imagination of men, that will command respect (or at least allegiance), and that will gradually assume world leadership. If Russia does this, eventually their way will be the way and we will have lost the struggle without knowing just when and how it occurred. The final victory in the struggle may well go to the country that offers mankind the greatest scientific achievements.37

 

The memo suggested that the NACA coordinate all American space-related research, both within the government and by industry. To have a successful space program, new knowledge had to be generated. Lundin took a shot at the "skeptical Congressmen and budget keepers." Even he, however, was not bold enough to suggest a development or a mission role for the new agency.38 Silverstein shared Lundin's enthusiasm for a strong leadership role for the NACA in space research. He set up an informal committee to begin to formulate plans for an additional laboratory to be devoted to space flight research.

On December 18, Dryden called a meeting of the directors and associate directors of the laboratories to discuss the future role of the NACA. In preparation for this meeting, Silverstein honed Lundin's ideas into a memo called "Lewis Laboratory Opinion of a Future Policy and Course of Action for the NACA." At Headquarters, when Dryden called on Henry Reid and Floyd Thompson of Langley, they showed little enthusiasm for the idea of taking on a large role in a national space effort. Smith DeFrance of Ames emphatically opposed a NACA space initiative because he feared that the NACA would lose its identity.39 Silverstein, last to speak, pulled the "Lewis Laboratory Opinion" from his briefcase. Not only did he argue for a central coordinating role for the NACA, but also he strongly advised that a new NACA space flight laboratory be authorized by Congress. Silverstein's strong advocacy swung the opinion of the meeting in favor of a new space agency. The ideas expressed in the "Lewis Laboratory Opinion" became the basis for NACA space policy, known as the "Dryden Plan."40

That evening, Dryden and James Doolittle, the Chairman of the NACA Main Committee, hosted the long-remembered "Young Turks Dinner" in the California Room of the Statler Hotel in Washington, D.C. The hosts intended to give the middle management of the three laboratories a chance to express their views on whether and how to redirect the goals of their venerable research organization. Silverstein selected Walter Olson, Eugene Manganiello, and Demarquis Wyatt to represent Lewis Laboratory. The sober Dryden gracefully accepted the barbs flung at him by some of his inebriated staff, who called him too cautious. He would encounter the same criticism after he testified in the House of Representatives that placing a man in a space capsule was like shooting a lady out of a cannon. John Stack of Langley Laboratory went so far as to call Dryden an "old fogey." The Young Turks of the three laboratories were enthusiastic about jumping into the space arena.41

Two days later the seven members of the Lewis Research Planning Council met to formalize the appointment of a special committee to plan the new space flight, laboratory recommended in the "Lewis Opinion" document. The members of the committee were Howard Childs, Chairman, Edgar M. Cortright, Robert E. English, Edmund R. Johash, Bernard Lubarsky, Phillip N. Miller, and Isidore Warshawsky. William Mickelson served as secretary. They were given two months to produce the plan. On February 10, 1958, they submitted "A Program for Expansion of NACA Research in Space Flight Technology." Not surprisingly, given that it was conceived at Lewis Laboratory, the document defined the mission of the new laboratory in terms of launch [161] vehicles - a stable of chemical, nuclear, and electric rockets. It would cost $380 million over a five-year period. The cost of expanding existing laboratories over five years would cost $55 million per year. Notably absent from this early planning document was a consideration of the mechanics of how missions would actually be carried out, and there was almost no mention of manned space flight. The document recommended that the NACA staff be increased from 8000 to 17,000 over a three-year period. It called for a budget increase from $80 million to $180 million.42

By NACA standards, this was a bold and visionary plan, but in hindsight it reveals the tenacious grip of the NACA's past practice. The new space laboratory would generate the knowledge, the technical know-how, necessary to make space flight missions possible. The development of the actual hardware to be sent into space and the operation of those missions would be left as it had in the past, to industry and to the military, respectively. There may not, however, have been a clear laboratory consensus on this last issue. Bruce Lundin later recalled that he "definitely thought that we should be responsible for the building and operation of launch vehicles and spacecraft just as we had always built and operated test models in our wind tunnels." In Lundin's view, the members of Silverstein's special committee represented the ideas of the group "that was afraid of getting 'sucked up' into space and who preferred the comfortable haven of their "research beds."43

The significance of the February 10 document in the evolving plans for the new space agency is not entirely clear. In 1965, when he was interviewed by Mercury historians Lloyd Swenson, James Grimwood, and Charles Alexander, Dryden recalled what is unmistakably the February 10 document as "Lewis Labs' bid for a lot of propulsion - nuclear propulsion - all sorts of things and a launch site:' Dryden criticized Lewis planners for neglecting manned space flight. He recalled his own past opposition, as well as the opposition of the von Braun group at Huntsville, Ala., to NACA participation in the "big rocket business."

 

Now their principle problem was that Huntsville would have nothing to do with them, and they did get into the rocket business on a small scale, but I always resisted their building up big facilities for Thors, and that sort of thing. So this [the plans expressed in the February 10 document] represented a sort of culmination of many efforts of the propulsion laboratory to get into rocket propulsion on a big scale, just as they were in engine - jet engine - propulsion. 44

 

Dryden claimed that the Lewis document had no status and never went further than his desk drawer. This is doubtful, since the document was later cited in the significant memo of March 5 from President Eisenhower as evidence of the NACA's competence to direct the new space agency.45

Initially it was not at all clear that the NACA would become the government agency around which the new National Aeronautics and Space Administration would be built. The external perceptions of the NACA after Sputnik, expressed both by the nation's scientific community and members of Congress, was that the NACA was too conservative to lead the exploration of space. Presumably, what made the NACA conservative in the eyes of the scientists was that it did applied research, an activity with a lower status than the work of the Navy's Vanguard group or university-based space science studies. To lawmakers, the NACA was a small agency without experience in the management of the huge budgets typical of defense contracts. The Cold War had accustomed Washington to think in terms of large-scale technology development. Firm centralized control of all three elements of the proposed space program - research, development, and [162] operations - seemed required to calm the nation's fears of Soviet technical superiority. In the end, the NACA won the leadership of the new space agency, not because of its positive qualities as a research organization, but to avoid having the new agency become an instrument for the militarization of space and greater Cold War competition. Eisenhower favored building NASA around the NACA because he wanted a civilian agency to direct the conquest of space. "Such reasoning," according to historian Walter McDougall, "made elevation of the innocuous NACA an attractive answer to the question of what to do about outer Space."46

There is no doubt that, as a small federal agency in the national security state of the 1950s, the NACA had taken a conservative approach to funding. Technically, however, the NACA could be quite daring. Dryden had encouraged the staff at all three laboratories to explore advanced technical concepts. National aeronautical policy had stipulated that fundamental research was the province of the NACA. Development was industry's prerogative. Operations belonged to the military.47 During the so called "lean years" between 1952 and 1957, the organizational structure of the NACA allowed its three laboratories Langley, Ames, and Lewis considerable autonomy. Many of the technical and scientific problems the laboratories tackled during these years belie the label of conservative. Obviously, each laboratory had to respond to the needs of the military and industry, but the NACA also prided itself on anticipating the technical needs of the nation by laying a base of knowledge on which to build future development.

Although it was understood that all NACA research would ultimately be applied to advance aircraft technology, projects whose immediate commercial applications were not clear always attracted considerable interest at the three NACA laboratories. Ames Laboratory's long and sustained interest in the problem of aerodynamic heating culminated in Harvey Allen's blunt body theory, later applied to the shape of the Mercury reentry capsule. In Engineer in Charge James Hansen describes the audacious reach of Langley's X-15 program to the edge of space.48 The X-series of research aircraft, a joint program with the Air Force and the Navy, involved building aircraft prototypes. As applied research, the program came close to the line separating research from development, but it yielded valuable technical and scientific data.

Three examples of NACA research at Lewis Laboratory belie the label of conservative: high-energy rocket fuels, nuclear rocket propulsion, and electric rocket propulsion. Technically, two of the three areas of research-nuclear rocket propulsion and electric propulsion - have yet to find applications. They were too visionary rather than too conservative. The third, liquid hydrogen, contributed substantially to the ability of the United States to land human beings on the Moon. Because of the momentum that these space-related projects had developed prior to the launching of Sputnik in October 1957, the transition to space was not as dramatic as the transition from the piston engine to jet propulsion. Lewis was already primed for space. Sputnik took the lid off the pent-up desire for more funding and recognition for areas that were already receiving considerable emphasis. As Business Week reported, Lewis Laboratory "leaped over to space" because it had begun its jump well before Sputnik.49 Nevertheless, in the context of the total research program, these efforts prior to Sputnik were small. Like the transition from the piston engine to jet propulsion, it was late. The NACA stayed too long in air-breathing engine technology. Silverstein missed the early opportunity to throw his full weight behind the efforts of the rocket group to develop high-energy rocket propellants.

 

 

FROM NACA TO NASA

 

[163] On March 5, 1959, President Eisenhower announced his decision to organize the new space agency around the NACA. In his memo justifying the decision in favor of a civilian agency, he cited three documents to underscore the interest and competence of the NACA for leadership of the space program: the so-called "Dryden Plan:' based on the ideas found in Silverstein's "Lewis Opinion," the February 10 plan for the new Space Flight Laboratory generated at Lewis, and the January 16 resolution of the Main Committee (drafted by the NACA's Special Committee on Space Technology, chaired by H. Guyford Stever). This resolution endorsed the role of the NACA as the national coordinator of space research.50

Between March and July, when the Space Act that authorized the National Aeronautics and Space Administration became law, a very different organization than the one envisioned by Lewis planners took shape. Although they had proposed a role for the NACA that they considered radical, because of the public furor raised in the wake of Sputnik, the new space agency could not be organized according to the old NACA model. After considerable debate within the Congress, the new agency was given extensive responsibility for development and operations (or missions), in addition to research.51 Moreover, the NACA had to give up its flexible committee structure and the autonomy of its laboratories in favor of a more centralized management. Although it is often stated that the NACA became the nucleus for NASA, this is not entirely correct, because the NACA purpose and way of doing business was entirely transformed.

Once the NACA's mandate was clear, Dryden asked Silverstein to assist him in the creation of the new space agency. Silverstein turned down Dryden's first invitation to relocate to Washington, D.C., but when asked a second time in spring 1958, he complied. Among the leaders of the NACA, Silverstein was the best suited to take on the job of pulling together the diverse elements of the new space flight development programs. He had drive and organizational ability. For the first three years of the agency, Silverstein's Office of Space Flight Programs had full responsibility for the Mercury Program and NASA's unmanned satellite programs. Silverstein named the Apollo program, and his office also laid the early groundwork for the manned lunar landing. Because of Silverstein's technical competence, the force of his personality, and the quality of the staff he selected to serve under him, power at Headquarters was concentrated in his hands until Headquarters reorganized in mid-1961.52

When Dryden was passed over in favor of T. Keith Glennan as NASA's first administrator, Glennan took a leave of absence from the Presidency of Case Institute of Technology. Having served as a Commissioner of the Atomic Energy Commission from 1950 to 1952 and on the Board of the National Science Foundation, he had good Washington connections. Glennan wisely insisted that Dryden remain as NASA's Deputy Administrator. Both he and Dryden were sworn in by President Eisenhower on August 19, 1958. He was catapulted from a job managing an operating budget at Case of $6 to 7 million, to one of $615 million, Glennan approved the NACA plans for the Mercury Program with an enthusiastic "Let's get going and don't spare the horses!"53

Silverstein had assembled an impressive group of former Lewis staff in his Office of Space Flight Programs. He was used to fostering innovation in engines at Lewis Laboratory. At Headquarters, his innovations would be administrative. Nevertheless, Silverstein's people had a feel for the technology they supervised. Demarquis Wyatt, a former Assistant Chief in charge of Lewis's supersonic wind tunnels, knew how to keep day-to-day operations running smoothly. After a few weeks in Washington, Silverstein asked Wyatt to become his Technical Assistant in charge of budgets, personnel, and trouble shooting. Wyatt faced a significant problem in figuring out how [164] to develop and manage a budget for the Mercury Program. Robert Gilruth, completely caught up in its technical demands, allowed financial details of payroll and contractor estimates to slide. Wyatt devised a financial control system for project management to streamline the unwieldy accounting system inherited from the NACA.

Silverstein selected Harold Finger, who had taken part in Lewis's "nuclear school," to become chief of NASA's nuclear programs. Francis C. Schwenk, also from Lewis, set to work under Finger to develop nuclear rocket concepts. Newell D. Sanders, who had first served under Silverstein in the Wind Tunnels and Flight Research Division, became Assistant Director for Advanced Technology. Edgar M. Cortright, Jr., and George Low, fraternity brothers at Rensselaer Polytechnic Institute, also took important positions in the fledgling agency. Both had come to Lewis Laboratory in 1949 with masters degrees in aeronautical engineering. Cortright, once Branch Chief for the 8-foot x 6-foot supersonic tunnel, had also taken part in the nuclear school. He had supervised some of the early planning to develop electric propulsion concepts. At Headquarters, Silverstein put Cortright in charge of Advanced Technology Programs in the Office of Space Flight. He headed NASA's meteorological satellite program, which included TIROS and Nimbus. Later he became Assistant Director for the Unmanned Lunar and Planetary Program.

George Low became Silverstein's deputy in charge of Manned Space Flight Programs. He shuttled tirelessly between Robert Gilruth at Langley and Silverstein to keep communications open and Mercury operations running smoothly. John Disher and Warren North, also spirited away from Lewis Laboratory, took key posts in Low's operation. They became early advocates of a lunar landing. In 1959, when Low served on NASA's Research Steering Committee on Manned Space Flight, chaired by Harry Goett, he pressed the committee to consider the Moon as a goal to follow after the Mercury program. By October 1960, Low had the green light from Silverstein to set in motion the first formal planning for a manned lunar program. Low's memo to Silverstein tersely stated, "It has become increasingly apparent that a preliminary program for manned lunar landings should be formulated. This is necessary in order to provide a proper justification for Apollo, and to place Apollo schedules and technical plans on a firmer foundation." Low advised the formation of a working group consisting of Oran Nicks, Eldon Hall, and John Disher. Silverstein's curt response, "O.K., Abe," hurriedly scrawled at the bottom of the memo, set in motion the careful technical planning that culminated nine years later in the "giant leap for mankind" - the historic lunar landing. 54

Silverstein also took charge of the initial planning of Goddard Space Flight Center. Working with speed and informality, Silverstein chose the site for Goddard and suggested its name. He played a major role in negotiations with the Navy to bring the Vanguard team to Goddard. Silverstein stepped in to serve as acting director of the Space Flight Center until its mission within NASA was sorted out and he was able to prevail upon Harry Goett from Ames to become its new director.

Like the generation from Langley in the early 1940s who shaped the Cleveland laboratory, the team from Lewis imprinted the new NASA organization with NACA traditions as they attempted to modify old ways in response to the new goals of NASA. The pervasive NACA influence was not always appreciated by new NASA employees recruited from outside the NACA. At Goddard, where the core staff came from the Naval Research Laboratory, a disgruntled employee circulated cards embossed with the message "Help Stamp Out NACA Types." Although NACA influence at Headquarters gradually diminished, during the first four years of the fledgling agency, Silverstein was the "lynchpin in the whole effort."55

 


[
165]

NACA's Special Committee on Space Technology, called the Stever Committee after its chairman, Guyford Stever, meets at Lewis, May 26, 1958.

NACA's Special Committee on Space Technology, called the Stever Committee after its chairman, Guyford Stever, meets at Lewis, May 26, 1958. Left to right, Edward R. Sharp, Director of Lewis Laboratory; Colonel Norman C Appold, US Air Force, Assistant to the Deputy Commander for Weapons Systems, Air Research and Development Command: Abraham Hyatt, Research and Analysis Officer, Bureau of Aeronautics, Department of the Navy; Hendrik W Bode, Director of Research Physical Sciences, Bell Telephone Laboratories; W Randolph Lovelace II, Lovelace Foundation for Medication Education and Research, S. K Hoffman, General Manager, Rocketdyne Division, North American Aviation; Milton U Clauser, Director, Aeronautical Research Laboratory, The Ramo-Wooldridge Corporation, H. Julian Allen, Chief, High Speed Flight Research, NACA Ames, Robert R. Gilruth, Assistant Director, NACA Langley, J. R. Dempsey, Manager. Convair-Astronautics; Carl B. Palmer, Secretary to Committee, NACA Headquarters, H. Guyford Stever, Chairman, Associate Dean of Engineering, Massachusetts Institute of Technology, Hugh L. Dryden (ex officio), Director, NACA; Dale R. Corson, Department of Physics, Cornell University; Abe Silverstein, Associate Director, NACA Lewis; Wernher von Braun, Director, Development Operations Division, Army Ballistic Missile Agency.

 

One of Glennan's chief objectives was to supplement the core NACA personnel and facilities by acquiring the cream of the nation's science and engineering talent in satellite and rocket development. In his view, although the NACA staff was "composed of reasonably able people," they lacked "experience in the management of large affairs."56 He met his first objective in December 1958, when the Army transferred to NASA its contract for the management of the Jet Propulsion Laboratory (JPL). After a faltering start, the JPL staff became NASA's experts in planetary and lunar probes.57

In Glennan's view, launch vehicles were the "limiting factor in the development of the nation's space program."58 The Mercury astronauts were lofted into space by rockets with payload capacities well below those of Russian rockets. Glennan knew that the only way to trump the Russians in the space race was to acquire the large rocket expertise of the von Braun team at the Army Ballistic Missile Agency in Huntsville, Ala. The Army, however, resisted giving up its [166] own plans for a manned space program. Threatened with termination of funds for the development of Huntsville's huge Saturn rocket, the Army at last capitulated in fall 1959.59

 

ENGINEERING, PLUMBING, AND PLAIN HOPE: SATURN V AND CENTAUR

 

The development of the Saturn rocket was a technical gamble because it went well beyond the existing rocket technology of both the United States and the Soviet Union. Glennan called it "one of the most amazing combinations of engineering, plumbing and plain hope that anyone could imagine."60 NASA policy makers thought that it was the key to leap-frogging the Russians' early successes in space.61 They were counting on a continuation of the strong ties with the American aerospace industry established by the missile program.

In December 1959 Silverstein assembled a committee to evaluate the proposed Saturn vehicle. The committee's specific charge was to decide on the configuration of the upper stages for Saturn, the vehicle that would power the astronauts to the Moon. Saturn had the potential to loft a heavy payload into space, but the configuration of its upper stages was a question with important implications. An upper stage fueled by liquid hydrogen could produce approximately 40 percent more payload per pound of lift-off weight than conventional propellants.

Silverstein's commitment to liquid hydrogen was a product of his experience with research on unconventional fuels at Lewis Laboratory. Project Bee had convinced Silverstein that hydrogen could be used safely. von Braun, however, did not share Silverstein's sanguine view of liquid....

 

 


T. Keith Glennan, the new NASA Administrator, and Hugh L. Dryden (left), Deputy Administrator, are sworn in August 19, 1958, as President Eisenhower looks on.

T. Keith Glennan, the new NASA Administrator, and Hugh L. Dryden (left), Deputy Administrator, are sworn in August 19, 1958, as President Eisenhower looks on.

 


[
167]

Senior staff and secretaries ring down the curtain on the NACA, September 30, 1958, in Director Ray Sharp's office.

Senior staff and secretaries ring down the curtain on the NACA, September 30, 1958, in Director Ray Sharp's office.

 

 

....hydrogen. Fearing risks involved in the storage and handling of liquid hydrogen, von Braun urged the use of conventional kerosene-based fuels for the first two stages, with liquid hydrogen reserved only for the third stage. Actually, it is doubtful that von Braun thought that liquid hydrogen would ever prove itself. In Silverstein's view,

 

The von Braun team was apparently willing to take on the difficulties of the 15 million-pound thrust-booster stage rather than the hazards which they contemplated in the use of hydrogen as a fuel. The Lewis team of NASA had pioneered in the use of hydrogen oxygen and had operated small rocket engines with hydrogen as a fuel for ten years prior to the time of the decision on Saturn. We had become very accustomed to its use and its safety.62

 

Silverstein came to the meeting of the Saturn Evaluation Committee armed with data from a NASA study on the feasibility of liquid hydrogen by Eldon Hall, Adelbert O. Tischler, and Abe Hyatt. He argued that conventional upper stages were too heavy for the required payload. For a week the NACA team and the von Braun group debated, until von Braun at last capitulated. The two upper stages of Saturn would use liquid hydrogen. "What led him to this final decision, I'll never know," Silverstein recalled, "but it was certainly the correct one... I believe that the decision to go with hydrogen-oxgyen in the upper stages of the Saturn V was the significant technical decision that enabled the United States to achieve the first manned lunar landing. The Russian effort to accomplish this mission without high-energy upper stages was doomed to failure."63

[168] Commitment to liquid hydrogen among rocket pioneers had an impressive history. Konstantin Tsiolkovsky recommended it in "Treatise on Space Ravel" in 1903. Hermann Oberth and Robert Goddard also discussed its potential. A group at Ohio State University tested a rocket using liquid hydrogen in 1945. By 1954, the rocket section at Lewis Laboratory had developed the nation's first regeneratively cooled liquid hydrogen-liquid fluorine rocket with 5000 pounds of thrust.64

Centaur originated in a proposal of Krafft Ehricke for using a liquid hydrogen upper stage for the Atlas missile developed by General Dynamics/Astronautics for the Air Force. Ehricke, once part of the Peenemunde group, had locked horns with von Braun over the practicality of liquid hydrogen in the face of the mind-boggling technical problems associated with using this unconventional fuel. The tragic explosion in the 1930s of the Hindenberg dirigible, lifted by gaseous hydrogen, had created an understandable prejudice against hydrogen in general. Rocket pioneers like von Braun had concrete reasons for doubting the feasibility of using liquid hydrogen as a fuel, Because hydrogen in its liquid state is cryogenic, it must be stored at very low temperatures. The temperature of liquid hydrogen at 1 atmosphere is -420° F. When it absorbs heat, it boils and expands rapidly. Tanks must be protected from all sources of heat, such as rocket engine exhaust, air friction during flight, and even radiant heat from the Sun. Because of liquid hydrogen's extreme cold and chemical reactivity with many metals, metal containers for hydrogen tend to become brittle and lose strength after exposure to hydrogen. Another serious drawback is that liquid hydrogen's low density requires light, but bulky, structures to contain the rocket's propellant. The von Braun group preferred heavy, solid structures - an approach Ehricke derisively compared to the conservative engineering of the Brooklyn Bridge. As John Sloop pointed out in Liquid Hydrogen as a Propulsion Fuel, "This conservative design philosophy mitigated against the use of liquid hydrogen which, more than conventional fuels, depended upon very light structures to help offset the handicap of low density."65

Ehricke, who had moved from Huntsville, Ala., to General Dynamics in San Diego, Calif., had tried to convince the Air Force to back his proposal for a liquid hydrogen upper stage on an Atlas intercontinental ballistic missile. The Air Force was not interested. In the post-Sputnik panic, Ehricke had more luck. He called on Silverstein in Washington in June 1958, during the transition from NACA to NASA. He wanted $15 million to initiate work at General Dynamics on a liquid hydrogen rocket. Silverstein shared Ehricke's enthusiasm for liquid hydrogen, but NASA had not yet received funding. He suggested that Ehricke present his proposal to the Advanced Research Projects Agency (ARPA).66

Shortly after NASA's formal creation in July 1958, Silverstein had set up a committee to define NASA's propulsion needs. By August this committee had concluded that, for launch vehicles requiring high-performance upper stages, the liquid hydrogen-liquid oxygen combination appeared to show the greatest promise. The consensus reached in this NASA committee coincided with ARMS acceptance of Ehricke's proposal. Pratt & Whitney, already versed in liquid hydrogen as a fuel for a gas turbine engine through its work on the Suntan project for the Air Force, agreed to undertake the development of the Centaur engine. Pratt & Whitney personnel had developed a good working relationship with Lewis engineers on this project when they visited Lewis several times to study injector designs for the RL-10 engine. One of the important steps in the development of this hydrogen engine was the use of regenerative cooling, long an interest of Lewis rocket researchers. Experimental heat transfer studies, carried out by a group working under Robert Graham in the Cryogenic Heat Transfer Section, proved the feasibility of hydrogen as a coolant.67 [169] Later, after Centaur management was transferred to Lewis Laboratory, Pratt & Whitney would work hand-in-hand with engineers at Lewis to complete the development of the RL-10 engine.

After the December 1959 meeting at which Silverstein had convinced von Braun that the upper stages of the Saturn rocket ought to be fueled with liquid hydrogen, the development of Centaur and Saturn V became inextricably intertwined. As an intermediate step NASA intended to use Centaur as the upper stage for an Atlas rocket to launch a probe to soft land on the Moon. No one was sure of the composition of the Moon's surface. Before astronauts could set foot there, NASA had to program this probe, named Surveyor, to land and photograph the Moon's surface. More important, the success of Centaur to launch Surveyor would prove the feasibility of using liquid hydrogen in Saturn's upper stages to power the astronauts to the Moon.

 

THE LUNAR DECISION AND NASA'S RESTRUCTURING

 

In May 1961 President John Kennedy announced the national goal of landing Americans on the Moon within a decade. Kennedy's commitment to a race with the Soviet Union transformed NASA. Its budget suddenly expanded from $1 billion in 1961 to a peak of $5.1 billion in 1964. The staff increased by a factor of ten, and Congress funded a new Manned Spaceflight Center in Houston, Tex. It was a bonanza for aerospace contractors and selected American universities, now able to develop new programs as a result of NASA's generous funding.

At Headquarters Silverstein had acted with the zest and informality necessary to get NASA off the ground. He had transformed an extremely capable and loyal team of engineers into effective NASA administrators. By 1961, Silverstein's Office of Space Flight Programs had become the operational hub of NASA. However, when James E. Webb became NASA Administrator in January 1961, he took steps to impose a more rational, hierarchical structure on Silverstein's seat-of-the-pants operation.

In June 1961 Webb set up a new Office of Programs under Demarquis Wyatt to report directly to Associate Administrator Robert Seamans. The creation of this office was designed to reduce the disproportionate concentration of power in Silverstein's Office of Space Flight Programs. 68 All the NASA centers would report directly to Seamans, a "frank recognition that the lunar landing decision had made manned spaceflight the dominant activity within the agency."69 Webb invited Silverstein to head the Apollo Program under the new centralized regime, but Silverstein resisted the organizational changes, arguing for a semi-autonomous status of the Apollo Program. He would agree to head the Apollo Program only if the centers directly responsible for the success of the program - Goddard, Marshall, the new Manned Space Flight Center in Houston, and Cape Canaveral - were placed directly under his supervision. In Silverstein's view, having the directors of these centers report to both the Apollo Program Director and to Associate Administrator Seamans would divide authority and generate misunderstandings and disagreement.70

One of the reasons Webb decided on this new management structure could have been the well-known friction between Silverstein and von Braun. With strong technical backgrounds and equally strong opinions, each had operated within separate spheres of NASA. Initially, when the Army Ballistic Missile Agency became Marshall Space Flight Center in March 1960, it was placed under the newly created Launch Vehicles Programs. However, if Silverstein were to direct the Apollo Program under the decentralized, semi-autonomous structure he favored, von Braun would become his subordinate, With von Braun and Silverstein "at loggerheads," the situation would have been intolerable.71 von Braun protested that he would not have his center run by a [170] "colony of artists" - his characterization of the NACA engineers associated with Silverstein in the Office of Space Flight Programs.72

On a more fundamental level, the September 1961 reorganization of Headquarters was an effort to cut NASA loose from the old NACA practice of decentralized administration. During his years as Director of the NACA, Dryden had served merely to coordinate the activity of the NACA laboratories. The locus of power, the real decision making, took place in the three laboratories, which received suggestions, not orders. The laboratories told Headquarters what "made sense or didn't make sense."73 Under the new regime, both technical and administrative decisions would originate from the top, in Washington.

Dryden urged Silverstein to take over the Apollo Program despite his misgivings. He reasoned that, once the management structure proved unworkable, it would be changed to the one Silverstein favored. To Silverstein this was a waste of time and money. Since he could not agree with the new policy, he felt he had no choice but to resign. With the helm of Lewis Laboratory unoccupied after the retirement of Ray Sharp the previous winter, Silverstein accepted the appointment. It was both an opportunity to return to research and a solution to his increasingly untenable position at Headquarters. Webb persuaded D. Brainerd Holmes from RCA, former manager of the Ballistic Missile Early Warning System, to head the Apollo Program.74

Abe Silverstein returned to Cleveland in November with the Mercury Program on the brink of success. Alan Shepherd's short flight beyond Earth's atmosphere had restored confidence in American technology. Now with a fellow Ohioan, John Glenn, scheduled to orbit Earth, Silverstein stepped out of the maelstrom of decision making at Headquarters to return to management and technical problems at the laboratory level.

 


Notes

1. See 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, New York, N.Y., 1963.

2. Ira H. Abbott, "A Review and Commentary of a Thesis by Arthur L. Levine Entitled, U.S. Aeronautical Research Policy 1915-1958: A Study of the Major Policy Decisions of the National Advisory Committee for Aeronautics' Dated 1963," HHN-35, 1964, NASA History Office, Washington, D.C., p. 197. Alex Roland has explored the conservatism of NACA Headquarters in Model Research, NASA SP-4103 (Washington, D.C.: U.S. Government Printing Office, 1985).

3. R. O. Miller and R M. Ordin, "Theoretical Performance of Rocket Propellants Containing Hydrogen, Nitrogen, and Oxygen," NACA RM E8A30, 1948, cited by John Stoop, Liquid Hydrogen as a Propulsion Fuel, 1949-1959, NASA SP-4404 (Washington, D.C.: U.S. Government Printing Office, 1978), p. 80.

4. Sloop, Liquid Hydrogen, p. 75.

5. Ibid.

6. R. W. Bussard and R.D.DeLauer, Fundamentals of Nuclear Flight (New York: McGraw Hill Book Co., 1965), p. 2.

7. John Sloop, Lectures 8 and 9 for ME 221, "Aircraft Propulsion Principles." Case Institute of Technology, fall, 1949. In Lecture 9, "Solid Propellant Rockets," Sloop discussed the range and orbital and escape velocities for satellites, p. 12-16. Private papers courtesy of John Sloop.

8. John Sloop, "Rocket Engines," American Legion Public Square Post, 12 June 1952, p. 6. Between 1948 and 1957 Sloop presented at least 15 papers at both professional and public meetings to urge rocket development for both missile and space applications. These papers, both published and unpublished, are available through Mr. Sloop, 8311 Melody Court, Bethesda, Md. 20034.

9. Sloop, Liquid Hydrogen, p. 78.

10. John Sloop, Comments on Manuscript History of the Lewis Research Center by Virginia Dawson, 5 September 1989.

11. On William Perl, see Ronald Radosh and Joyce Milton, The Rosenberg File: A Search for the Truth (New York: Holt, Rinehart and Winston, 1983), p. 123-129, 202-207, and passim. I also consulted the Perl Files (65-59312) at the FBI Headquarters, J. Edgar Hoover Building, Washington, D.C. Although the files revealed nothing beyond the inconclusive evidence of Perl's guilt reported in Radosh and Milton, the interviews are interesting for their descriptions of the laboratory's work during that period. Much of the text of the files is still heavily censored.

12. Memorandum for Director, NACA, from E. R. Sharp, "Visit of Congressman Albert Thomas and Mr. A. H. Skarin, Friday, Oct 3, 1952." NASA Lewis Records, 279/116.1-63.

13. Interview with John Evvard, 14 June 1987.

14. Notes from the Noncredit Graduate Study Course (1957), "Concepts of Hypersonic Flight," were obtained courtesy of George M. Prok, NASA Lewis Research Center.

15. Moeckel to Stuhlinger, 2 June 1964, personal papers of Moeckel.

16. Ernst Stuhlinger, Ion Propulsion for Space Flight (New York: McGraw-Hill Book Co., 1964), p. 6.

17. Moeckel to Stuhlinger, ibid.

18. K. A. Ehricke, "The Solar Powered Space Ship," ARS Paper No. 310-356.

19. Stuhlinger, Ion Propulsion for Space flight, p. 6.

20. Moeckel to Stuhlinger, 2 June 1964, personal papers of Moeckel.

21. Walter Olson, "A Suggested Policy and Course of Action for NACA with Regard to Rocket Engine Propulsion," 6 May 1955, p. 21-22 of "A Suggested Policy and Course of Action for NACA on Space Flight," 2 December 1957, part of communication, to Eugene Emme from Walter Olson, 31 January 1972, Propulsion file, NASA History Office, Washington, D.C.

22. See Abe Silverstein and Eldon W. Hall, "Liquid Hydrogen as a Jet Fuel for High-Altitude Aircraft," NACA RM E55 C28a, 15 April 1955. See also Sloop, Liquid Hydrogen, p. 102-107.

23. Interview with Robert English by V. Dawson, 11 July 1986. William Ritter's group (including Merland Moseson, Robert Ziemer, and Wilfred Scull) transferred to Goddard Space Flight Center.

24. Sloop, Liquid Hydrogen, p. 84.

25. Quoted by Sloop, Liquid Hydrogen, p. 85.

26. Interview with Robert English, 11 July 1986. Unfortunately, I was unable to find the curriculum of the nuclear school among what has survived of Lewis administrative records. Members of the Nuclear School were (list courtesy of Robert English):

From the Compressor and Turbine Division

Turbines: Hubert Allen, Art Hansen, Howard Herzig

Fundamental Turbines: R. Cavicchi, E. Davison, Robert English

Compressors: Dan Bernatowicz, Ted Fessler, Harry Finger, Sy Lieblein, Carl Schwenk

Turbine Cooling: Tony Diaguila, Pat Donoughe, Herman Ellerbrock, John Livingood, Henry Slone

From other divisions: Edgar Cortright, Tom Dallas, Irving Goodman, Eldon Hall, Jim Kramer, Herb Heppler, Bill Phillips, Lou Rosenblum

27. Syllabus of Seminar on Rocket Propulsion, 28 February 1957; Memorandum for Personnel Taking Rocket Seminar and for Instructional Staff, 4 March 1957. Papers of John Sloop.

28. Moeckel to Stuhlinger, 2 June 1964, personal papers of Moeckel. See Wolfgang Moeckel, L. V. Baldwin, Robert English, Bernard Lubarsky, and Steve Maslen, "Satellite and Space Propulsion Systems," NASA TN D-285, June 1960. Declassified paper presented at NACA Flight Propulsion Conference, 22 November 1957, previously published as NASA TMX 61622. Some of the Lewis work in the early 1960s is summarized in Electric Propulsion for Spacecraft, NASA SP-22, 1963.

29. Research Planning Council (established 8 March 1957), 6 September 1957, NASA Lewis Records 298/117.

30. "10 Years After X-1: Supersonic Anniversary Puts Flight Laboratory Here in Spotlight," Cleveland Plain Dealer Magazine, 6 October 1957.

31. Sloop, Liquid Hydrogen, p. 90-91.

32. Triennial Inspection, Lewis Laboratory, 7-9 October 1957. "High-Energy Rocket Propellants," personal papers courtesy of John Sloop.

33. Papers of the NACA 1957 Flight Propulsion Conference were published in two parts. NASA TM-X-67368 (1971) contains six papers on air-breathing propulsion systems, three on rocket systems, including "Propellants," by E. A. Fletcher, H. W, Douglass, R. J. Priem, and G. Vasu; "Turbopumps for High-Energy Propellants," by Ambrose Ginsburg, Ward W. Wilcox, and David G. Evans; and "Performance and Missions," by J. L. Sloop, A. S. Boksenvom, S. Gordon, R. W. Graham, R M. Ordin, and A. 0. Tischler. The papers published as TM-X-61622 (1972) include the paper cited in note 28; Paul G. Johnson, James W. Miser, and Roger L. Smith, "Nuclear Logistic Carrier," and Frank E. Rom, Eldon W. Sams, and Robert E. Hyland, "Nuclear Rockets."

34. Interview with Bruce Lundin by Eugene Emme, 5 June 1974. See discussion in Sloop, Liquid Hydrogen, p. 180.

35. Walter Olson, "Suggested Policy and Course of Action for NACA on Space Flight," 2 December 1957, part of a communication to Eugene Emme from Walter Olson, 31 January 1972. Propulsion file, NASA History Office, Washington, D.C.

36. A copy of the original memo, dated 9 December 1957, was obtained from Lundin, with Silverstein's hand written comments. The typed Silverstein version called, "Lewis Laboratory Opinion of a Future Policy and Course of Action for the NACA," was found in NASA Lewis Records, 221/115.1 71,

37. Ibid.

38. Ibid.

39. Sloop, Liquid Hydrogen, p. 180.

40. "Minutes of December 20, 1957 Meeting of Committee on Space Flight Laboratory," personal papers of Isidore Warshawsky. From these minutes it appears that the "Lewis Laboratory Opinion" was the original document that, with modifications and input from the other two laboratories, became "A Staff Study of the NACA," dated 14, January 1958. This was further refined into a four page document called "A National Research Program for Space Technology," referred to by later commentators as the "Dryden Plan."

41. See Sloop, Liquid Hydrogen, p. 181; Hugh L. Dryden to Eugene Emme, "The NACA-NASA Transition," 8 September 1965, NASA History Office, Washington, D.C.

42. Silverstein attended meetings on January 6 and 27 of the Lewis Special Committee on Space Flight Laboratory. From the personal papers of I. Warshawsky, I obtained the complete minutes of the meetings. The 10 February document is reproduced in toto by Alex Roland in Model Research, vol. 2, document 46, p. 732 ff. The final version, somewhat condensed and called "Summary of a Program for Expansion of NACA Research in Space Flight Technology," was found in NASA Lewis Records 295/117.71. From the 27 January 1958 minutes, it is clear that Langley and Ames offered similar proposals, but Headquarters decided to use the Lewis proposal for a brochure to be presented to the NACA Special Committee on Space Technology (Stever Committee). The final version is also attributed to Lewis senior staff by L. Swenson, Jr., et al., This New Ocean, NASA SP-4201 (Washington, D.C.: U.S. Government Printing Office, 1966), p. 76 and 533, note 7. This is presumably the earliest plan for God dard, a laboratory completely different from the original Lewis conception.

43. Letter from Bruce Lundin to V Dawson, 27 May 1987.

44. "Dryden Interview by Mercury Historians," 1 October 1965, Dryden file, Johnson Space Center Archives, Houston, Tex.

45. Robert L. Rosholt, An Administrative History of NASA, 1958#1963, NASA SP-4101, (Washington, D.C.: U.S. Government Printing Office, 1966).

46. Walter A. McDougall, . . . the Heavens and the Earth: A Political History of the Space Age (New York: Basic Books, 19851, p. 166. See also Homer E. Newell, Beyond the Atmosphere, NASA SP-4211 (Washington, D.C.: U.S. Government Printing Office, 19801 p. 90 91.

47. For the postwar statement of "National Aeronautical Research Policy," approved 21 March 1946, see Alex Roland, Model Research, vol. 2, document 36, p. 693-695. For distinctions between fundamental (or basic), applied, and specific developmental research, see Ira Abbott, p. 5-6 and 186 (cited in note 2). Arthur L. Levine (cited in note 11 placed the so-called "fundamental or basic research" at about 10 percent of total NACA research.

48. See James R. Hansen, Engineer in Charge, NASA SP 4305 (Washington, D.C.: U.S. Government Printing Office, 1987), p. 353 367.

49. "How an Aircraft Lab Leaped Over to Space," Business Week, 4 June 1960, p. 98-102.

50. Robert Rosholt, An Administrative History of NASA, p. 35. Although set up in November 1957, the Stever Committee did not actually meet until January 1958. It seems that by this time most of the NACA policy had been formulated, and the committee merely ratified already formulated NACA plans at that time.

51. For the metamorphosis of the NACA into NASA, see especially Rosholt, An Administrative History of NASA, p. 37-70; Swenson, This New Ocean, p. 75-106.

52. Rosholt, An Administrative History of NASA, p. 206-207. See also transcript of D. D. Wyatt interview with Eugene Emme, 21 June 1973, NASA History Office.

From Lewis Laboratory Silverstein selected Demarquis D. Wyatt, Assistant to the Director of Space Flight Development; Edgar M. Cortright, Chief, Advanced Technology Program; Harold B. Finger, Chief, Nuclear Engines Program; Eldon W. Hall, Chief, Analysis and Requirements Pro gram; N. Philip Miller, Chief, Plant and Facilities Construction Program, Space Flight Program. Also George M. Low, Chief, Manned Space Flight Program; Warren J. North, Chief, Manned Satellite Program; Adelbert O. Tischler, Chief, Liquid Fuel Rocket Engines Program; Francis C. Schwenk, Nuclear Program (title unknown); Newell Sanders, Assistant Director, Advanced Technology Program; William Fleming, Chief, Project Review Division (chaired task group to prepare the "Fleming Report," the first study of the feasibility of manned flight to the Moon).

Of the approximately 40 members of the Space Task Group (which became Mercury Program), 10 were from Lewis Laboratory: Elmer Buller, A. M. Busch, W. R. Dennis, M. J. Krasnican, Glynn S. Lunney, Andre J. Meyer, W. R. Meyer, W. J. Nesbitt, Gerard J. Pesman, and Leonard Robb. Commuters to Langley from Lewis were John Disher and Kenneth Weston.

53. Transcript of speech by Abe Silverstein, Kennedy Space Center Writers Conference, Cocoa Beach, Fla., 1-3 September 1977. NASA History Office.

54. Memo from George Low to Abe Silverstein, 17 October 1960, Personal files of Abe Silverstein, also reproduced in Charles Murray and Catherine Cox, Apollo: The Race to the Moon (New York: Simon and Schuster, 1989), p. 57-58. Their discussion confirms John Logsdon's view that, by 1959, at least two years prior to Kennedy's dramatic announcement of the Apollo Program, policy committees within NASA had set a manned lunar landing as NASA's long-range goal. See John M. Logsdon, The Decision to Go to the Moon:, Project Apollo and the National Interest (Cambridge, Mass.: MIT Press, 1970), p. 56-57; see also discussion in Roger Bilstein, Stages to Saturn: A Technological History of the Apollo/Saturn Launch Vehicles, NASA SP-4206 (Washington, D.C.: U.S. Government Printing Office, 1980), p. 48-50. See also Richard J. Weber to Chief, Propulsion Systems Analysis Branch, "Visit of R. J. Weber to Headquarters on Feb. 17, 1960 for ABMA briefing on lunar projects," files of R. J. Weber, NASA Lewis Research Center.

55. Wyatt interview with Emme, p. 99.

56. T. Keith Glennan, "The First Years of the National Aeronautics and Space Administration," 1964, unpublished diary, vol. 1, p. 6-7. Eisenhower Library, Abilene, Kan.

57. See Clayton R. Koppes, The JPL and the American Space Program (New Haven: Yale University Press, 1982). See also remarks of T Keith Glennan, "The First Years," p. 17.

58. T. Keith Glennan, "The First Years," Vol. 1, p. 18. For a history of Project Mercury, see Swenson, This New Ocean.

59. Herbert York, Making Weapons, Talking Peace (New York: Basic Books, 1987), p. 174-175.

60. T. Keith Glennan, "The First Years," vol. 2, p. 15.

61. Logsdon, The Decision to Go to the Moon, p. 117.

62. Abe Silverstein, "How It All Began," Speech at Kennedy Space Center Writers Conference, Cocoa Beach, Fla., 1-3 September 1977, NASA History Office Archives, file marked Lewis Research Center.

63. Abe Silverstein, "How It All Began!' See also Sloop, Liquid Hydrogen, p. 232-235, and Logsdon, The Decision to Go to the Moon, p. 58. Members of the Saturn Evaluation Committee were Silverstein, Chairman; Abraham Hyatt, NASA; George P. Sutton, ARPA; T. C. Muse, ODDR&E; Norman C. Appold, U.S. Air Force; Wernher von Braun, ABMA; Eldon Hall, NASA, Secretary. For key documents, see "Report to the Administrator, NASA on Saturn Development Plan by Saturn Vehicle Team," 15 December 1959, NASA History Office, Silverstein biographical file.

64. Sloop, Liquid Hydrogen, p. 91-93.

65. Ibid, p. 208,

66. Ibid, p. 191-197. Interview with Abe Silverstein by V. Dawson, 6 June 1987.

67. RVO classic papers on hydrogen heat transfer that grew out of the work with Pratt & Whitney are Robert C. Hendricks, Robert W Graham, Yih-yun Hsu, and Robert Friedman, "Experimental Heat Transfer and Pressure Drop of Liquid Hydrogen Flowing Through a Heated Tube," NASA TN D 765, May 1961; "Experimental Heat Transfer Results for Cryogenic Hydrogen Flowing in Tubes at Subcritical and Supercritical Pressures to 800 psi Absolute," NASA TN D 3095, March 1966.

68. Rosholt, An Administrative History of NASA, p. 206-207.

69. Arnold S. Levine, Managing NASA in the Apollo Era, NASA SP-4102 (Washington, D.C.: U.S. Government Printing Office, 1982), p. 36.

70. Interview with Abe Silverstein, 6 June 1987, not recorded.

71. Transcript of D. D. Wyatt interview with Eugene Emme, p. 96.

72. Interview with Bruce Lundin by V. Dawson, 28 May 1987.

73. Transcript of Wyatt interview with Emme, p, 68-70.

74. Silverstein interview with V. Dawson, 6 June 1987. Silverstein said that in November 1963 Webb asked him to head the proposed Manned Spacecraft Center in Houston, but he felt that Robert Gilruth had earned that. Charles Murray and Catherine Cox, Apollo, note p. 126, report that Silverstein opposed Gilruth's appointment. Dryden's prediction that the new management structure would ultimately be changed to the one Silverstein favored proved correct. The centers were later placed under program offices, instead of reporting to Associate Administrator Seamans. After Sharp retired in December 1960, Eugene Manganiello served as Acting Director until Silverstein returned.


previousindexnext