User:Overthrows/sandbox

From Wikisource
Jump to navigation Jump to search

my sacred testing zone how i love you so

Transclusion tests[edit]

I


The Lure, the Lock, the Key

(to 1958)

THE yearning of men to escape the confines of their Earth and to travel to the heavens is older than the history of mankind itself. Religion, mythology, and literature reaching back thousands of years are sprinkled with references to magic carpets, flying horses, flaming aerial chariots, and winged gods.[1] Although "science fiction" is a descriptive term of recent vintage, the fictional literature of space travel dates at least from the second century A.D. Around the year 160 the Greek savant Lucian of Samosata wrote satirically about an imaginary journey to the Moon, "a great countrie in the aire, like to a shining island," as Elizabethan scholars translated his description 1500 years later. Carried to the Moon by a giant waterspout, Menippus, Lucian's hero, returns to Earth in an equally distinctive manner: The angry gods simply have Mercury take hold of his right ear and deposit him on the ground. Lucian established a tradition of space-travel fiction, and generations of later storytellers spawned numerous fantasies in which by some miraculous means—such as a flight of wild lunar swans in a seventeenth-century tale by Francis Godwin or a cannon shot in Jules Verne's classic account of a Moon voyage (1865-1870)—earthlings are transported beyond the confines of their world and into space.[2]

But apparently the first suggestion, fictional or otherwise, for an artificial manned satellite of Earth is to be found in a short novel called "The Brick Moon," written in 1869 by the American Edward Everett Hale and originally serialized in the Atlantic Monthly. Although, like most of his contemporaries, Hale had only a vague notion of where Earth's atmosphere ended and where space began, he did realize that somewhere the "aire" became the "aether," and he also understood the mechanics of putting a satellite into an Earth orbit:

If from the surface of the earth, by a gigantic peashooter, you could shoot a pea upward …; if you drove it so fast and far that when its power of ascent was exhausted, and it should fall, it should clear the earth …; if you had given it sufficient power to get it half way round the earth without touching, that pea would clear the earth forever. It would continue to rotate… with the impulse with which it had first cleared our atmosphere and attraction.

Dare’s mid-19th century illustration, "A Voyage to the Moon," captured man's age-old dream of lifting himself off Earth and venturing out toward our celestial neighbors, the Moon, the Sun, the planets, and even the stars.


The action of centripetal forces as advanced by Isaac Newton: "That by means of centripetal forces the planets may be retained in certain orbits, we may easily understand, if we consider the motions of projectiles; for a stone that is projected is by the pressure of its own weight forced out of the rectilinear path, which by the initial projection alone it should have pursued, and made to describe a curved line in the air; and through that crooked way is at last brought down to the ground; and the greater the velocity is with which it is projected, the farther it goes before it falls to the earth. We may therefore suppose the velocity to be so increased, that it would describe an arc of 1,2, 5, 10, 100, 1,000 miles before it arrived at earth, till at last, exceeding the limits of the earth, it should pass into space without touching it." In Hale’s story a group of industrious New Englanders construct a 200-foot-diameter brick sphere, which, carrying 37 people, is prematurely hurled into an orbit 4000 miles from Earth by two huge flywheels.[3] Less than a hundred years later, Hale's own country would undertake a more modest and more practicable scheme for a manned satellite in Project Mercury.

Centuries before Hale wrote about an orbiting manned sphere, Nicolaus Copernicus, Johannes Kepler, Galileo Galilei, and other astronomers had helped put the solar system in order, with the Sun in the center and the various planets, spherical and of different sizes, orbiting eclliptically around it. Isaac Newton had established the basic principles of gravitation and mechanics governing reaction propulsion and spatial navigation.[4] Thus it was possible for Hale and his fellow fictionists to think at least half seriously about, and to describe in fairly accurate detail, such adventures as orbiting Earth and its Moon and voyaging to Venus.

Most flight enthusiasts in the nineteenth century, however, were absorbed with the problems of flight within the atmosphere, with conveyance from one place to another on Earth. This preoccupation with atmospheric transport, which would continue until the mid-twentieth century, in many ways retarded interest in rocketry and space travel. But the development and refinement of aeronautics in the twentieth century was both a product of and a stimulant to man’s determination to fly ever higher and faster, to travel as far from his Earth as he could. Atmospheric flight, in terms of both motivation and technology, was a necessary prelude to the exploration of near and outer space. In a sense, therefore, man's journey along the highway to space, leading to such astronautical achievements as Project Mercury, began in the dense forest of his atmosphere, with feats in aeronautics.


Conquest of the Air

Man first ventured aloft in balloons in the 1780s, and in the next century gliders also bore human passengers on the air. By 1900 a host of theoreticians and inventors in Europe and the United States were steadily expanding their knowledge and capability beyond the flying of balloons and gliders and into the complexities of machineborne flight. The essentials of the airplane—wings, rudders, engine, and propeller—already were well known, but what had not been done was to balance and steer a heavier-than-air flying machine.

On December 8, 1903, Samuel Pierpont Langley, a renowned astrophysicist and Secretary of the Smithsonian Institution, tried for the second time to fly his manned "aerodrome," a glider fitted with a small internal combustion engine, by catapulting it from a houseboat on the Potomac River. The much-publicized experiment, financed largely by the United States War Department, ended in failure when the machine plunged, with pilot-engineer Charles M. Manley, into the cold water.[5] The undeserved wave of ridicule and charges of waste that followed Langley’s failure obscured what happened nine days later at Kitty-Hawk, North Carolina. There two erstwhile bicycle mechanics from Dayton, Ohio, Wilbur and Orville Wright, carried out "the first [flight] in the history of the world in which a machine carrying a man had raised itself by its own power into the air in full flight, had sailed forward without reduction of speed, and had finally landed at a point as high as that from which it started."[6] Although few people realized it at the time, practicable heavier-than-air flight had become a reality.

The United States Army purchased the first military airplane, a Wright Flyer, in 1908. But when Europe plunged into general war in 1914, competitive nationalism—drawing on the talents of scientists like Ernst Mach in Vienna, Ludwig Prandtl in Germany, and Osborne Reynolds in Great Britain, and of inventors like the Frenchmen Louis Bleriot and Gabriel Voisin—had accelerated European flight technology well beyond that of the United States.[7] In 1915, after several years of agitation for a Government-financed "national aeronautical laboratory" like those already set up in the major European countries, Congress took the first step to regain the leadership in aeronautics that the United States had lost after 1908. By an amendment attached to a naval appropriation bill, Congress established an Advisory Committee for Aeronautics "to supervise and direct the scientific study of the problems of flight, with a view to their practical solution." President Woodrow Wilson, who at first had feared that the creation of such an organization might reflect on official American neutrality, appointed the stipulated 12 unsalaried members to the "Main Committee," as the policymaking body of the new organization came to be called. At its first meeting, the Main Committee changed the name of the organization to National Advisory Committee for Aeronautics, and shortly "NACA" began making surveys of the state of aeronautical research and facilities in the country. During the First World War it aided significantly in the formulation of national policy on such critical problems as the cross-licensing of patents and aircraft production. NACA did not have its own research facilities, however, until 1920, when it opened the Langley Memorial Aeronautical Laboratory, named after the "aerodrome" pioneer, at Langley Field, Virginia.[8]

In the 1920s and 1930s aeronautical science and aviation technology continued to advance, as the various cross-country flights, around-the-world flights, and the most celebrated of all aerial voyages, Charles A. Lindbergh’s nonstop flight in 1927 from New York to Paris, demonstrated. During these decades NACA brought the United States worldwide leadership in aeronautical science. Concentrating its research in aerodynamics and aerodynamic loads, with lesser attention to structural materials and powerplants, NACA worked closely with the Army and Navy laboratories, with the National Bureau of Standards, and with the young and struggling aircraft industry to enlarge the theory and technology of flight.[9] The reputation for originality and thorough research that NACA quietly built in the interwar period would continue to grow until 1958, when the organization would metamorphose into a glamorous new space agency, the likes of which might have frightened the early NACA stalwarts.

Early
Aviation

Man's liberation from the surface of Earth began at Kitty Hawk, N.C., on December 17, 1903, when Orville and Wilbur Wright made the world’s first controlled, powered flights in a heavier-than-air machine (above). At last it was within man's grasp to use Earth’s atmosphere as a means of transportation. There was much to learn; in the United States the National Advisory Committee for Aeronautics pioneered aeronautical research in the 1920s. Early wind tunnel research at Langley Memorial Aeronautical Laboratory (right) culminated in the famous NACA cowling and the family of NACA wing shapes that would dominate several generations of aircraft from the 1920s into the 1940s. And aviation finally came of age in world opinion with the epochal solo flight from New York to Paris by Charles E. Lindbergh, May 20–21, 1927. Lindbergh and his plane, the "Spirit of St. Louis," arc shown (below, right) visiting the Washington Navy Yard on June 11, 1927.

Over the years NACA acquired a highly competent staff of "research engineers" and technicians at its Langley laboratory.[10] Young aeronautical and mechanical engineers just leaving college were drawn to NACA by the intellectual independence characterizing the agency, by the opportunity to do important work and see their names on regularly published technical papers, and by the superior wind tunnels and other research equipment increasingly available at the Virginia site. NACA experimenters made discoveries leading to such major innovations in aircraft design as the smooth cowling for radial engines, wing fillets to cut down on wing-fuselage interference, engine nacelles mounted in the wings of multi-engine craft, and retractable landing gear. This and other research led to the continual reduction of aerodynamic drag on aircraft shapes and consequent increases in speed and overall performance.[11]

The steady improvement of aircraft design and performance benefited commercial as well as military aviation. Airlines for passenger, mail, and freight transport, established in the previous decade both in the United States and Europe, expanded rapidly in the depression years of the thirties. In the year 1937 more than a million passengers flew on airlines in the United States alone.[12] At the same time, advances in speed, altitude, and distance, together with numerous innovations in flight engineering and instrumentation, presaged the arrival of the airplane as a decisive military weapon.[13]

Yet NACA remained small and inconspicuous; as late as the summer of 1939 its total complement was 523 people, of whom only 278 were engaged in research activities. Its budget for that fiscal year was $4,600,000.[14] The prevailing mood of the American public throughout the thirties was reflected in the neutrality legislation passed in the last half of the decade, in niggardly defense appropriations, and in the preoccupation of the Roosevelt administration with the domestic aspects of the Great Depression. Without greatly increased appropriations from Congress, the military was held back in its efforts to acquire more and better aerial weapons. Without a military market for its products, the American aircraft industry proceeded cautiously and slowly in the design and manufacture of airframes and powerplants. And in the face of the restricted needs of industry and the armed services and severly limited appropriations, NACA kept its efforts focused where it could acquire the greatest quantity of knowledge for the smallest expenditure of funds and manpower—in aerodynamics.

As Europe moved nearer to war, however, the Roosevelt administration, Congress, and the public at large showed more interest in an expanded military establishment, including military aviation. Leading figures like Lindbergh and Vannevar Bush, president of the Carnegie Institution and chairman of the Main Committee, warned of the remarkable gains in aviation being made in other countries, especially in Nazi Germany.[15] While the United States may have retained its aerodynamics research lead, the Germans, drawing, in part from the published findings of NACA, by 1939 had temporarily outstripped this country in aeronautical development.

After the outbreak of war in Europe, NACA eventually secured authorization and funding to increase its program across the board, including a much enlarged effort in propulsion and structural materials research. A new aeronautical laboratory, named after physicist Joseph S. Ames of Johns Hopkins University, former chairman of the Alain Committee, was constructed beginningin 1940 on land adjacent to the Navy installation at Moffett Field, California, 40 miles south of San Francisco. The next year, on a site next to the municipal airport at Cleveland, NACA broke ground for still another laboratory, to be devoted primarily to engine research. In later years the Cleveland facility would be named the Lewis Flight Propulsion Laboratory, after George W. Lewis, for 28 years NACA’s Director of Research.[16]

Some nine months before Pearl Harbor, Chairman Bush of NACA appointed a Special Committee on Jet Propulsion, headed by former Main Committeeman William F. Durand of Stanford University, and including such leaders in aeronautical science as Theodore von Karman of the California Institute of Technology and Hugh L. Dryden of the National Bureau of Standards.[17] Until then NACA, the military services, and the aircraft industry had given little attention to jet propulsion. There had been little active disagreement with the conclusion reached in 1923 by Edgar Buckingham of the Bureau of Standards: "Propulsion by the reaction of a simple jet cannot complete, in any respect, with air screw propulsion at such flying speeds as are now in prospect."[18] By 1941, however, Germany had flown turbojets, and her researchers were working intensively on the development of an operational jet-propelled interceptor. In Britain the propulsion scientist Frank Whittle had designed and built a gas-turbine engine and had flown a turbojet-powered aircraft.

Faced with the prospect of European-developed aircraft that could reach flight regimes in excess of 400 miles per hour and operational altitudes of about 40,000 feet, NACA gradually authorized more and more research on jet powerplants for the Army Air Forces and the Navy. Most of the NACA research effort during the war, however, went to “quick fixes,” improving or "cleaning up" military aircraft already produced by aircraft companies, rather than to the more fundamental problems of aircraft design, construction, and propulsion.[19] So, understandably and predictably, during the Second World War, Germany was first to put into operation military aircraft driven by jet powerplants, as well as rocket-powered interceptors that could fly at 590 miles per hour and climb to 40,000 feet in two and a half minutes.[20] The German jets and rocket planes came into the war too late to have any effect on its outcome, but the new aircraft caused consternation among American aeronautical scientists and military planners.

The Second World War saw, in the words of NACA Chairman Jerome C. Hunsaker, "the end to the development of the airplane as conceived by Wilbur and Orville Wright."[21] Propeller-driven aircraft advanced far beyond their original reconnaissance and tactical uses and became integral instruments of strategic warfare. The development of the atomic bomb meant a multifold increase in the firepower of aircraft, but well before the single B—29 dropped the single five-ton bomb on Hiroshima, long-range bomber fleets carrying conventional TNT explosives and incendiaries had radically altered the nature of war.[22]

The frantic race in military technology developing in the postwar years between the United States and the Soviet Union produced a remarkable acceleration in the evolution of the airplane. Jet-propelled interceptors, increasingly rakish in appearance by comparison with their staid propeller-driven ancestors, flew ever faster, higher, and farther.[23] Following the recommendations of a series of blue-ribbon scientific advisory groups, the Defense Department and the newly independent Air Force made the Strategic Air Command, with its thousands of huge manned bombers, the first line of American defense in the late forties and early fifties.[24] To many people the intercontinental bomber, carrying fission and (after 1954) hydrogen-fusion weapons, capable of circumnavigating the globe nonstop with mid-air refueling, looked like the “ultimate weapon” men had sought since the beginning of human conflict.

Working under the incessant demands of the cold-war years, NACA continued to pioneer in applied aeronautical research. By 1946 the NACA staff had grown to about 6800, its annual budget was in the vicinity of $40 million, and its facilities were valued at more than $200 million. Although Chairman Hunsaker and others on the Main Committee felt that NACA's principal mission should be inquiry into the fundamentals of aeronautics, the military services and the aircraft industry continued to rely on NACA as a problem-solving agency. The pressure for "quick fixes" persisted as the Korean War intensified requirements for work on specific aircraft problems.[25]

The outstanding general impediment to aeronautical progress, however, continued to be the so-called "sonic barrier," a region near the speed of sound (approximately 750 miles per hour at sea level, 660 miles per hour above 40,000 feet) wherein an aircraft encounters compressibility phenomena in fluid dynamics, or the "piling up" of air molecules. A serious technical obstacle to high-speed research in the postwar years was the choking effect experienced in wind tunnels during attempts to simulate flight conditions in the transonic range (600-800 miles per hour). A wind tunnel constructed at Langley employing the slotted-throat principle to overcome the choking phenomenon did not begin operation until 1951, and a series of NACA and Air Force supersonic tunnels, authorized by Congress under the Unitary Plan Act of 1949, was not completed until the mid-fifties.[26] NACA investigators had to use other methods for extensive transonic research. One was a falling-body technique, in which airplane models equipped with radio-telemetry apparatus were dropped from bombers at high altitudes. Another was the firing of small solid-propellant rockets to gather data on various aerodynamic shapes accelerated past mach, the speed of sound. Many of these tests supported military missile studies. The rocket firings were carried out at the Pilotless Aircraft Research Station, a facility set up by the Langley laboratory on Wallops Island, off the Virginia coast, in the spring of 1945. The Flightless Aircraft Research Division at Langley, until the early fifties headed by Robert R. Gilruth, conducted the NACA program of aerodynamic research with rocket-launched models.[27]

The most celebrated part of the postwar aeronautical research effort in the United States, however, was the NACA-military work with rocket-propelled aircraft. In 1943, Langley aerodynamicist John Stack and Robert J. Woods of the Bell Aircraft Corporation, realizing that propeller-driven aircraft had about reached their performance limits, suggested the development of a special airplane for research in the problems of transonic and supersonic flight. The next year, the Army Air Forces, the Navy, and NACA inaugurated a program for the construction and operation of such an airplane, to be propelled by a liquid-fueled rocket engine. Built by Bell and eventually known as the X–1, the plane was powered by a 6000-pound-thrust rocket burning liquid oxygen and a mixture of alcohol and distilled water. On October 14, 1947, above Edwards Air Force Base in southern California, the X—1 dropped from the underside of its B–29 carrier plane at 35,000 feet and began climbing. A few seconds later the pilot of the small, bullet-shaped craft, Air Force Captain Charles E. Yeager, became the first man officially to fly faster than the speed of sound in level or climbing flight.[28]

The X–1 was the first of a line of generally successful rocket research airplanes. In November 1953 the Navy’s D-558-II, built by the Douglas Aircraft Company and piloted by A. Scott Crossfield of NACA, broke mach 2, twice sonic speed; but this record stood only until the next month, when Yeager flew the new Bell X–1A to mach 2.5, or approximately 1612 miles per hour. The following summer Major Arthur Murray of the Air Force pushed the X–1A to a new altitude record of 90,000 feet above the Mojave Desert test complex consisting of Edwards Air Force Base and NACA’s High Speed Flight Station. These spectacular research flights, besides banishing the myth that aircraft could not fly past the “sonic barrier,” affected the design and performance of tactical military aircraft.[29] In the early fifties, the Air Force and the aircraft industry, profiting from the mountain of NACA research data, were preparing to inaugurate the new “century series” of supersonic jet interceptors.[30] And representatives of NACA, the Air Force, and the Navy Bureau of Aeronautics already were planning a new experimental rocket plane, the X–15, to employ the most powerful rocket aircraft motor ever developed and to fly to an altitude of 50 miles, the very edge of space.

Thus less than a decade after the end of the Second World War, airplanes—jet-powered and rocket-propelled—had virtually finished exploring the sensible atmosphere, the region below 80,000 or 90,000 feet. Much work remained for aeronautical scientists and engineers in such areas as airflow, turbulence, engines, and fuels, but researchers in NACA, the military, and the aircraft industry approached the thorniest problems in aeronautics with a confidence grounded in 50 years of progress. Man’s facility in atmospheric flight and his adjustment to the airplane seemed complete. Pilots had mastered some of the most complex moving machines ever contrived, and passengers sat comfortably and safely in

The famous research aircraft series is shown above: in the center, the Douglas X–3; lower left, the Bell X–1A; continuing left to right, the Douglas D-558-I, the Convair XF–92A, the Bell X–5, the Douglas D–558–II, and the Northrop X–4. In the photo below, the X–15 is shown as it drops away from its mother B–52 and starts its own 57,000-lb.-thrust engine to begin another of its highly successful research flights.

pressurized cabins on high-altitude airliners featuring an unprecedented combination of speed and luxury. It appeared that man at last had accomplished what the ancients had dreamed of—conquest of the air.


The Highway to Space

Space flight, however, was something else. While in one sense atmospheric flight was the first step toward space flight, extra-atmospheric transport involves much more than a logical extension of aviation technology. The airplane, powered either by a reciprocating or a jet engine, is a creature and a captive of the atmosphere, because either pbwerplant depends on air—more properly, oxygen—for its operation, and in space there is no air. But the rocket, unlike the gas turbine, pulsejet, ramjet, or piston engine, needs no air. It carries everything needed for propulsion within itself—its own fuel and some form of oxidizer, commonly liquid oxygen, to burn the fuel. So the rocket engine operates independently of its environment; in fact, its efficiency increases as it climbs away from the frictional density of the lower atmosphere to the thin air of the stratosphere and into the airlessness of space.[31]

Yet even the rocket research airplanes were a long way from spacecraft. Although some of these vehicles provided data on the use of reaction controls for steering in the near vacuum of the upper atmosphere, they were designed to produce considerable aerodynamic lift for control within the lower atmosphere; and, in terms of the mass to be accelerated, their powerplants burned too briefly and produced too little thrust to counterbalance the oppressive force of gravity. Fulfillment of the age-old desire to travel to the heavens, even realization of Hale’s nineteenth century concept of a manned sphere circling Earth in lower space, would have to await the development of rockets big enough to boost thousands of pounds and to break the lock of gravity.

Although black-powder rockets, invented by the Chinese, had been used for centuries for festive and military purposes, not until the late nineteenth and early twentieth centuries did imaginative individuals in various parts of the world begin seriously to consider the liquid-fueled rocket as a vehicle for spatial conveyance. The history of liquid-fueled rocketry, and thus of manned space flight, is closely linked to the pioneering careers of three men—the Russian Konstantin Eduardovich Tsiolkovsky (1857–1935), the American Robert Hutchings Goddard (1882–1945), and the German-Romanian Hermann Oberth (1894–1989).

Tsiolkovsky, for most of his life an obscure teacher of mathematics, authored a series of remarkable technical essays on such subjects as reaction propulsion with liquid-propellant rockets, attainable velocities, fuel compositions, and oxygen supply and air purification for space travelers. He also wrote what apparently was the first technical discussion of an artificial Earth satellite.[32] Although virtually unknown in the West at the time of his death, in 1935, Tsiolkovsky was honored by the Soviets and had helped establish a long Russian tradition of astronautics. This tradition helps to account for the U.S.S.R.'s advances with rocket-assisted airplane takeoffs and small meteorological rockets of the 1930s and her space achievements of the 1950s and 1960s.[33]

In terms of experimentation, Goddard, professor of physics at Clark University, was by far the most important of the rocket pioneers.[34] As early as 1914 he secured a patent for a small liquid-fueled rocket engine. Six years later he published a highly technical paper on the potential uses of a rocket with such an engine for studying atmospheric conditions at altitudes from 20 to 50 miles. Toward the end of the paper he mentioned the possibility of firing a rocket containing a powder charge that could be exploded on the Moon. "It remains only to perform certain necessary preliminary experiments before an apparatus can be constructed that will carry recording instruments to any desired altitude," he concluded.[35]

Goddard's life for the next 20 years was devoted to making those "necessary preliminary experiments." Working in the 1920s in Massachusetts with financial support from various sources and in the New Mexico desert with Guggenheim Foundation funds during the succeeding decade, Goddard compiled an amazing list of "firsts" in rocketry. Among other things, he carried out the first recorded launching of a liquid propellant rocket (March 16, 1926), adapted the gyroscope to guide rockets, installed movable deflector vanes in a rocket exhaust nozzle for stability and steering, patented a design for a multistage rocket, developed fuel pumps for liquid-rocket motors, experimented with self-cooling and variable-thrust motors, and developed automatically deployed parachutes for recovering his instrumented rockets. Finally, he was the first of the early rocket enthusiasts to go beyond theory and design into the realm of "systems engineering"—the complex and hand-dirtying business of making airframes, fuel pumps, valves, and guidance devices compatible, and of doing all the other things necessary to make a rocket fly. Goddard put rocket theory into practice, as his 214 patents attest.[36]

Goddard clearly deserves the fame that has attached to his name in recent years, but in many was he was more inventor than scientist. He deliberately worked in lonely obscurity, jealously patented virtually all of his innovations, and usually refused to share his findings with others. Consequently his work was not as valuable as it might have been to such of his contemporaries as the young rocket buffs who formed the American Rocket Society in the early thirties and vainly sought his counsel.[37]

Goddard’s disdain for team research prompted his refusal to work with the California Institute of Technology Rocket Research Project, instigated in 1936 by the renowned von Kármán, then director of the Guggenheim Aeronautical Laboratory at CalTech. The CalTech group undertook research in the fundamentals of high-altitude sounding rockets, including thermodynamics, the principles of reaction, fuels, thrust measurements, and nozzle shapes. Beginning in 1939 the Guggenheim Laboratory, under the first Federal contract for rocket

From theory through laboratory demonstration through design, construction, test flight, and use of payload, Robert H. Goddard must rank as the U.S. pioneer in modern rocketry. The famous photo at the right shows Goddard beside his first successful liquid-fuel rocket, flown March 16, 1926. Years later, in the spring of 1941, he had progressed to larger, more complex models, like the one shown below in his workshop at Mescalero Ranch, Roswell, N. Mex., with his assistants. In December 1944, Goddard sent this photo to his long-time benefactor Harry F. Guggenheim with the comment, "It is practically identical with the German V–2 rocket."

research, carried out studies and experiments for the Army Air Forces, especially on rocket-assisted takeoffs for aircraft. These takeoff rockets were called JATO (for "Jet-Assisted Take-Off") units, because, as one of the CalTech scientists recalled, "the word 'rocket' was of such bad repute that [we] felt it advisable to drop the use of the word. It did not return to our vocabulary until several years later …"[38] In 1944, with the Guggenheim Laboratory working intently on Army and Navy contracts for JATO units and small bombardment rockets, the Rocket Research Project was reorganized as the Jet Propulsion Laboratory.[39]

In the 1920s and 1930s interest in rocketry and space exploration became firmly rooted in Europe, although the rapid expansion of aviation technology occupied the attention of most flight-minded Europeans. Societies of rocket theorists and experimenters, mostly privately sponsored, were established in several European countries.[40] The most important of these groups was the Society for Space Travel (Verein für Raumschiffahrt), founded in Germany but having members in other countries. The "VfR," as its founders called it, gained much of its impetus from the writings of Oberth, who in 1923, as a young mathematician, published his classic treatise on space travel, The Rocket into Interplanetary Space. A substantial portion of this small book was devoted to a detailed description of the mechanics of putting into orbit a satellite of Earth.[41]

Spurred by Oberth's theoretical arguments, the Germans in the VfR in the early thirties conducted numerous static firings of rocket engines and launched a number of small rockets. Meanwhile the German Army, on the assumption that rocketry could become an extension of long-range artillery and because the construction of rockets was not prohibited by the Treaty of Versailles, had inaugurated a modest rocket development program in 1931, employing several of the VfR members. One of these was a 21-year-old engineer named Wernher von Braun, who later became the civilian head of the army’s rocket research group. In 1933 the new Nazi regime placed all rocket experimentation, including that being done by the rest of the VfR, under strict government control.[42]

The story of German achievements in military rocketry during the late thirties and early forties at Peenemuende, the vast military research installation on the Baltic Sea, is well known.[43] Knowing Goddard's work only through his published findings, the German experimenters contrived and elaborated on nearly all of the American’s patented technical innovations, including gyroscopic controls, parachutes for rocket recovery, and movable deflector vanes in the exhaust. The rocket specialists at Peenemuende were trying to create the first large, long-range military rocket. By 1943, after numerous frustrations, they had their "big rocket," 46 feet long by 11½ feet in diameter, weighing 34,000 pounds when fueled, and producing 69,100 pounds of thrust from a single engine consuming liquid oxygen and a mixture of alcohol and water. Called "Assembly–4" (A–4) by the Peenemuende group, the rocket had a range of nearly 200 miles and a maximum velocity of about 3500 miles per hour, and was controlled by its gyroscope and exhaust deflector vanes, sometimes supplemented by radio control.[44] When Major General Walter Domberger, commander of the army works at Peenemuende, pronounced the A–4 operational in 1944, Joseph Goebbels' propaganda machine christened it Vergeltungswaffe zwei (Vengeance Weapon No. 2), or "V–2."[45] But for the space-travel devotees at Peenemuende the rocket remained the A–4, a step in the climb toward space.

Although the total military effect of the 3745 V–2s fired at targets on the Continent and in England was slight, this supersonic ballistic missile threw a long shadow over the future of human society. As the Western Allies and the Soviets swept into Germany, they both sought to confiscate the elements of the German rocket program in the form of records, hardware, and people. Peenemuende was within the Russian zone of occupation, but before the arrival of the Soviet forces von Braun and most of the other engineers and technicians fled westward with a portion of their technical data. The Americans also captured the underground V–2 factory in the Harz Mountains; 100 partially assembled V–2s were quickly dismantled and sent to the United States. Ultimately von Braun and about 125 other German rocket specialists reached this country under "Project Paperclip," carried out by the United States Army.[46]

The Soviets captured no more than a handful of top Peenemuende engineers and administrators. "This is absolutely intolerable," protested Josef Stalin to

Hermann Oberth with key officials of the Army Ballistic Missile Agen- cy at Huntsville, Ala., in 1956. Counterclockwise from the left: Maj. Gen. H. N. T of toy, com- manding general of ABMA, who organized Project Paperclip; Ernst Stuhlinger; Oberth; Wern- her von Braun, Director, Develop- ment Operations Division; and Eberharcl Rees, Deputy Director, Development Operations Division.

17 Lieutenant Colonel G. A. Tokaty, one of his rocket experts. "We defeated the Nazi armies; we occupied Berlin and Peenemuende; but the Americans got the rocket engineers."[47] The Russians did obtain a windfall, however, in the form of hundreds of technicians and rank-and-file engineers, the Peenemuende laboratories and assembly plant, and lists of component suppliers. From those suppliers located in the Russian zone the Soviets secured enough parts to reactivate the manufacture of V–2s. The captured technicians and engineers were transported to the Soviet Union, where the Russian rocket specialists systematically drained them of the technical information they possessed but did not permit them to participate directly in the burgeoning postwar Soviet rocket development program.[48]

During the war Russian rocket developers, like their American counterparts, had concentrated on JATO and small bombardment rockets. "Backward though they were often said to be in matters of technology," observed James Phinney Baxter right after the war, "it was the Russians who in 1941 first employed rockets on a major scale. They achieved a notable success, and made more use of the rocket as a ground-to-ground weapon than any other combatant."[49] In the postwar years the Soviets quickly turned to the development of large liquid-propellant rockets. Lacking an armada of intercontinental bombers carrying atomic warheads, such as the United States possessed, they envisioned "trans-Atlantic rockets" as "an effective straight jacket for that noisy shopkeeper Harry Truman," to use Stalin’s words.[50] Consequently the U.S.S.R. undertook to build a long-range military rocket years before nuclear weaponry actually became practicable for rockets; indeed, even before the Soviets had perfected an atomic device for delivery by aircraft.

The U.S.S.R. began exploration of the upper atmosphere with captured V–2s in the fall of 1947. Within two years, however, Soviet production was underway on a single-stage rocket called the T–1, an improved version of the V–2. The first rocket divisions of the Soviet Armed Forces were instituted in 1950 or 1951. Probably in 1954, development work began on a multistage rocket to be used both as a weapon and as a vehicle for space exploration. And in the spring of 1956 Communist Party Chairman Nikita Khrushchev warned that "soon" Russian rockets carrying thermonuclear warheads would be able to hit any target on Earth.[51]


Postwar American Rocketry

Meanwhile the United States, convinced of the long-term superiority of her intercontinental bombers, pursued national security by means of airpower. The extremely heavy weight of atomic warheads meant that they would have to be delivered by large bombers, or by a much bigger rocket than anyone in the military was willing to ask Congress to fund. Despite the early postwar warnings of General Henry H. Arnold and others, for whom the V–2 experience was prophetic, the Truman administration and Congress listened to conservative military men and civilian scientists who felt that until at least 1965 manned bombers, supplemented by air-breathing guided missiles evolving from the German V–1, should be the principal American "deterrent force."[52] Just after the war former NACA Chairman Bush, then Director of the Office of Scientific Research and Development, had expressed the prevailing mood in a much-quoted (and perhaps much- regretted) piece of testimony before a Congressional committee: "There has been a great deal said about a 3000-mile high-angle rocket. In my opinion, such a thing is impossible today and will be impossible for many years… I wish the American public would leave that out of their thinking."[53]

The United States developed guided missiles for air-to-air, air-to-surface, and surface-to-air interception uses and as tactical surface-to-surface weapons. Rocket motors, using both liquid and solid fuels, gradually replaced jet propulsion systems, but short-range defensive missiles remained advanced enough for most tastes until the late 1950s.[54]

As for scientific research in the upper atmosphere, the backlog of V–2s put together by the United States Army from captured components would do in the early postwar years. From April 1946 to October 1951 , 66 V–2s were fired at the Army's White Sands Proving Grounds, New Mexico, in the most extensive rocket and upper-atmospheric research program to that time. The Army Ordnance Department, the Air Force, the Air Force Cambridge Research Center, the General Electric Company, various scientific institutions, universities, and government agencies, and the Naval Research Laboratory participated in the White Sands V–2 program. Virtually all the rockets were heavily instrumented, and many of them carried plant life and animals. V–2s carried monkeys aloft on four occasions; telemetry data transmitted from the rockets showed no ill effects on the primates until each was killed in the crash. The most memorable launching at White Sands, however, came on February 24, 1949, when a V–2 boosted a WAC Corporal rocket developed by the Jet Propulsion Laboratory 244 miles into space and to a speed of 5510 miles per hour, the greatest altitude and velocity yet attained by a man-made object. A year and a half later, a V–2—WAC Corporal combination rose from Cape Canaveral, Florida, in the first launch at the Air Force’s newly activated Long Range Proving Ground.[55]

By the late forties, with the supply of V–2s rapidly disappearing, work had begun on more reliable and efficient research rockets. The most durable of these indigenous projectiles proved to be the Aerobee, designed as a sounding rocket by the Applied Physics Laboratory of Johns Hopkins University and financed by the Office of Naval Research. With a peak altitude of about 80 miles, the Aerobee served as a reliable tool for upper-atmospheric research until the late 1950s.[56] The Naval Research Laboratory designed the Viking, a long, slim high-altitude sounding rocket, manufactured by the Glenn L. Martin Company of Baltimore. In August 1951 the Viking bettered its own altitude record for a single-stage rocket, reaching 136 miles from a White Sands launch. In the fifties,

Launch of the record-setting U.S. Army-Jet Propulsion Laboratory Bumper WAC (V-2 first stage and WAC/Corporal second stage) from White Sands Proving Ground, N. Mex. The first Bumper-WAC launch occurred on May 13, 1948. On February 24, 1949, the two-stage rocket reached its record altitude of 244 miles and speed of 5150 miles per hour.


instrumentation carried in Aerobees and Vikings extended knowledge of the atmosphere to 150 miles, provided photographs of Earth's curvature and cloud cover, and gave some information on the Sun and cosmic radiation.[57]

In 1955 the Viking was chosen as the first stage and an improved Aerobee as the second stage for a new; three-stage rocket to be used in Project Vanguard, which was to orbit an instrumented research satellite as part of the American contribution to the International Geophysical Year. The decision to use the Viking and the "Aerobee-Hi" in this country’s first effort to launch an unmanned scientific satellite illustrates the basic dichotomy in thought and practice governing postwar rocket development in the United States: After the expenditure of the V–2s, scientific activity should employ relatively inexpensive sounding rockets with small thrusts. Larger, higher-thrust, and more expensive rockets to be used as space launchers must await a specific military requirement. Such a policy meant that the Soviet Union, early fostering the ballistic missile as an intercontinental delivery system, might have a proven long-range rocket before the United States; the Soviets might also, if they chose, launch larger satellites sooner than this country.

By 1951, three sizable military rockets were under development in the United States. One, an Air Force project for an intercontinental ramjet-booster rocket combination called the Navaho, took many twists and turns before ending in mid-1957. After 11 years and $680 million, the Air Force, lacking funds for further development, canceled the Navaho enterprise. Technologically, however, Navaho proved a worthwhile investment; its booster-engine configuration, for example, became the basic design later used in various rockets.[58] The two other rocket projects being financed by the military in the early fifties were ultimately successful, both as weapons systems and as space boosters.


Redstone and Atlas

After the creation of a separate Air Force in 1947, the Army had continued rocket development, operating on the same assumption behind the German Army's research in the 1930s—that rocketry was basically an extension of artillery. In June 1950, Army Ordnance moved its team of 130 German rocket scientists and engineers from Fort Bliss at El Paso to the Army's Redstone Arsenal at Huntsville, Alabama, along with some 800 military and General Electric employees. Headed by Wernher von Braun, who later became chief of the Guided Missile Development Division at Redstone Arsenal, the Army group began design studies on a liquid-fueled battlefield missile called the Hermes C1, a modified V–2. Soon the Huntsville engineers changed the design of the Hermes, which had been planned for a 500-mile range, to a 200-mile rocket capable of high mobility for field deployment. The Rocketdyne Division of North American Aviation modified the Navaho booster engine for the new weapon, and in 1952 the Army bombardment rocket was officially named "Redstone."[59]

Always the favorite of the von Braun group working for the Army, the Redstone was a direct descendant of the V–2. The Redstone's liquid-fueled engine burned alcohol and liquid oxygen and produced about 75,000 pounds of thrust. Nearly 70 feet long and slightly under 6 feet in diameter, the battlefield missile had a speed at burnout, the point of propellant exhaustion, of 3800 miles per hour. For guidance it utilized an all-inertial system featuring a gyroscopically stabilized platform, computers, a programmed flight path taped into the rocket before launch, and the activation of the steering mechanism by signals in flight. For control during powered ascent the Redstone depended on tail fins with movable rudders and refractory carbon vanes mounted in the rocket exhaust. The prime contract for the manufacture of Redstone test rockets went to the Chrysler Corporation. In August 1953 a Redstone fabricated at the Huntsville arsenal made a partially successful maiden flight of only 8000 yards from the military’s missile range at Cape Canaveral, Florida. During the next five years, 37 Redstones were fired to test structure, engine performance, guidance and control, tracking, and telemetry.[60]

The second successful military rocket being developed in 1951 was an Air Force project, the Atlas. The long history of the Atlas, the first American continental ballistic missile (ICBM),[61] in began early in 1946, when the Air Materiel Command of the Army Air Forces awarded a study contract for a long-range missile to Consolidated Vultee Aircraft Corporation (Convair), of San Diego. By mid-year a team of Convair engineers, headed by Karel J. Bossart, had completed a design for "a sort of Americanized V–2," called “HIROC,” or Project MX-774. Bossart and associates proposed a technique basically new to American rocketry (although patented by Goddard and tried on some German V–2s)—controlling the rocket by swiveling the engines, using hydraulic actuators responding to commands from the autopilot and gyroscope. This technique was the precursor of the gimbaled engine method employed to control the Atlas and other later rockets. In 1947, the Truman administration and the equally economy-minded Republican 80th Congress confronted the Air Force with the choice of having funds slashed for its intercontinental manned bombers and interceptors or cutting back on some of its advanced weapons designs. Just as the first MX-774 test vehicle was nearing completion, the Air Force notified Convair that the project was canceled. The Convair engineers used the remainder of their contract funds for static firings at Point Loma, California, and for three partially successful test launches at White Sands, the last on December 2, 1948.[62]

From 1947 until early 1951 there was no American project for an intercontinental ballistic missile. The Soviet Union exploded her first atomic device in 1949, ending the United States' postwar monopoly on nuclear weapons. President Harry S. Truman quickly ordered the development of hydrogen-fusion warheads on a priority basis. The coming of the war in Korea the next year shook American self-confidence still further. The economy program instituted by Secretary of Defense Louis Johnson ended, and the military budget, including appropriations for weapons research, zoomed upward. The Army began its work leading to the Redstone, while the Air Force resumed its efforts to develop an intercontinental military rocket. In January 1951 the Air Materiel Command awarded Convair a new contract for Project MX-1593, to which Karel Bossart and his engineering group gave the name "Project Atlas."[63] Yet the pace of the military rocket program remained deliberate, its funding conservative.

A series of events beginning in late 1952 altered this cautious approach. On November 1, at Eniwetok Atoll in the Pacific, the Atomic Energy Commission detonated the world's first thermonuclear explosion, the harbinger of the hydrogen bomb. The device weighed about 60,000 pounds, certainly a much greater weight than was practicable for a ballistic missile payload. The next year, however, as a result of a recommendation by a Department of Defense study group, Trevor Gardner, assistant to the Secretary of the Air Force, set up a Strategic Missiles Evaluation Committee to investigate the status of Air Force long-range missiles. The committee, composed of nuclear scientists and missile experts, was headed by the famous mathematician John von Neumann. Specifically, Gardner asked the committee to make a prediction regarding weight as opposed to yield in nuclear payloads for some six or seven years hence. The evaluation group, familiarly known as the "Teapot Committee," concluded that shortly it would be possible to build smaller, lighter, and more powerful hydrogen-fusion warheads. This in turn would make it possible to reduce the size of rocket nose cones and propellant loads and, with a vastly greater yield from the thermonuclear explosion, to eliminate the need for precise missile accuracy.[64] In February 1954 both the Strategic Missiles Evaluation Committee and the Rand Corporation, the Air Force-sponsored research agency, submitted formal reports predicting smaller nuclear warheads and urging that the Air Force give its highest priority to work on long-range ballistic missiles.

Between 1945 and 1953 the yield of heavy fission weapons had increased substantially from the 20-kiloton bomb dropped on Hiroshima. Now, according to the Air Force’s scientific advisers, lighter, more compact, and much more powerful hydrogen warheads could soon be realized. These judgments "completely changed the picture regarding the ballistic missile," explained General Bernard A. Schriever, who later came to head the Air Force ballistic missile development program, "because from then on we could consider a relatively low weight package for payload purposes."[65] This was the fateful "thermonuclear breakthrough."

Late in March 1964 the Air Research and Development Command organized a special missile command agency, originally called the Western Development Division but renamed Air Force Ballistic Missile Division on June 1, 1957. Its first headquarters was in Inglewood, California; its first commander, Brigadier General Schriever. The Convair big rocket project gained new life in the winter of 1954-55, when the Western Development Division awarded its first long-term contract for fabrication of an ICBM. The awarding of the contract came in an atmosphere of mounting crisis and urgency. The Soviets had exploded their own thermonuclear device in 1953, and intelligence data from various sources indicated that they also were working on ICBMs to carry uranium and hydrogen warheads. Thus the Atlas project became a highest-priority “crash” program, with the Air Force and its contractors and subcontractors working against the fearsome possibility of thermonuclear blackmail.[66]

Rejecting the Army-arsenal concept, whereby research and development and some fabrication took place in Government facilities, the Air Force left the great bulk of the engineering task to Convair and its associate contractors.[67] For close technical and administrative direction the Air Force turned to the newly formed Ramo-Wooldridge Corporation, a private missile research firm, which established a subsidiary initially called the Guided Missiles Research Division, later Space Technology Laboratories (STL). With headquarters in Los Angeles, the firm was to oversee the systems engineering of the Air Force ICBM program.[68]

In November 1955, STL's directional responsibilities broadened to include work on a new Air Force rocket, the intermediate-range (1800-mile) Thor, hastily designed by the Douglas Aircraft Company to serve as a stopgap nuclear deterrent until the intercontinental Atlas became operational. At the same time

Military missiles of the 1950s provided both the technology and the first-generation boosters for the nascent space program. The Air Force's Navaho (left) was a long-range cruise missile overtaken by the onrush of technology; though it was canceled as a project, it had pioneered the development of large rocket engines and guidance systems. The Atlas missile (center) had a hectic on-and-off career in the early 1950s but became the first operational ICBM and the major "large" boost vehicle for manned and unmanned space missions in the first decade of the space age. Thor (right), the sturdy, reliable baby of the Atlas technology, served an interim military role as an operational IRBM and a longer and more illustrious role as the workhorse booster of the first decade of payloads for military and nonmilitary space projects. Shown here with an Able second stage, it accepted a variety of second stages and payloads. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/40 Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/41 Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/42 Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/43 Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/44 Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/45 Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/46 Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/47 Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/48

  1. See Gertrude and James Jobes, Outer Space: Myths, Name Calendars, Meanings: From the Emergence of History to the Present Day (New York, 1964).
  2. For the long history of space travel fiction see Marjorie Hope Nicolson, Voyages to the Moon (paperback ed., New York, 1960), and Science and Imagination (Ithaca, N.Y., 1965); Willy Ley, Rockets, Missiles, and Space Travel (Rev. ed., New York, 1957), 9-40; Arthur C. Clarke, "Space Travel in Fact and Fiction," Journal of the British Interplanetary Society, IX (Sept. 1950), 213-230; James O. Bailey, Pilgrims Through Space and Time: Trends and Patterns in Scientific and Utopian Fiction (New York, 1957); Roger L. Green, Into Other Worlds: Space-Flight from Lucian to Lewis (New York, 1958); Philip B. Gove, The Imaginary Voyage in Prose and Fiction: A History of Its Criticism and a Guide to Its Study… (New York, 1961); John Lear, Kepler's Dream (Berkeley, Calif., 1965); and W. R. Maxwell, "Some Aspects of the Origins and Early Development of Astronautics," Journal of the British Interplanetary Society, XVIII (Sept. 1962), 415-425.
  3. Edward Everett Hale, "The Brick Moon," Atlantic Monthly, XXIV (Oct., Nov., Dec., 1869), 451-460, 603-611, 679-688. Also published in Hale, The Brick Moon and Other Stories (New York, 1899). Hale is of course better known for another story, "The Man Without a Country."
  4. Good treatments of astronomical developments in the 16th, 17th, and 18th centuries are in A. R. Hall, The Scientific Revolution, 1500-1800: The Foundation of the Modern Scientific Attitude (Boston, 1954); and Alexandre Koyre, From the Closed World to the Infinite Universe (paperback ed., New York, 1958).
  5. Charles G. Abbot, Great Inventions (Washington, 1943), 227-229. On Langley's failure and the public reaction to it, see Mark Sullivan, Our Times: The United States, 1900-1925, Vol. II: America Finding Herself (New York, 1927), 562-564. In 1914, after numerous modifications and largely as an attempt to invalidate the Wright Brothers' patents, Glen H. Curtiss flew the Langley aerodrome successfully with pontoons. Fourteen years later the Smithsonian reconciled itself to the fact the Wrights’ airplane of 1903 was the first successful flying machine, rather than Langley’s aerodrome. See Abbot, "The Relations between the Smithsonian Institution and the Wright Brothers," Smithsonian Miscellaneous Collections, LXXXI (Sept. 29, 1928).
  6. Orville Wright, quoted in N. H. Randers-Pehrson, History of Aviation (New York, 1944), 36. For a description of the flight, see Elsbeth E. Freudenthal, Flight into History: The Wright Brothers and the Air Age (Norman, Okla., 1949), 3-90; Marvin W. McFarland, ed., The Papers of Wilbur and Orville Wright… (2 vols., New York, 1953), I, 395-397; and Charles H. Gibbs-Smith, "The Wright Brothers and Their Invention of the Practical Aeroplane," Nature, CXCVIII (June 1, 1963), 824-826.
  7. There are several reasonably good histories of aviation and aeronautical research, including M. J. B. Davy, Interpretive History of Flight (London, 1948); Charles H. Gibbs-Smith, The History of Flying (New York, 1954) and The Aeroplane (London, 1960); Lloyd Morris and Kendall Smith, Ceiling Unlimited: The Story of American Aviation from Kitty Hawk to Supersonics (New York, 1953) ; Theodore von Karman, Aerodynamics: Selected Topics in the Light of Their Historical Development (Ithaca, N.Y., 1954); and R. Giacomelli, "Historical Sketch," in William F. Durand, cd., Aerodynamic Theory: A General Review of Progress (2 ed., 6 vols. in 3, New York, 1963), I, 304-394. See also Hunter Rouse and Simon Ince, History of Hydraulics (Iowa City, Iowa, paperback ed., New York, 1963), 229-242.
  8. Jerome C. Hunsaker, "Forty Years of Aeronautical Research," Report of the Smithsonian Institution for 1955 (Washington, 1956), 241-251; Arthur S. Levine, "United States Aeronautical Research Policy, 1915—1958: A Study of the Major Policy Decisions of the National Advisory Committee for Aeronautics,” unpublished Ph. D. dissertation, Columbia University, 1963, 7-16; George W. Gray, Frontiers of Flight: The Story of NACA Research (New York, 1948), 9-15; A. Hunter Dupree, Science in the Federal Government: A History of Policies and Activities to 1940 (Cambridge, Mass., 1957), 283-287; John F. Victory, "The NACA: Cradle of Research," Flying, LX (March 1957), 40-43. In 1921, NACA installed at Langley a pioneering variable-density wind tunnel, which featured the use of compressed air to produce an airflow over small models, thus closely simulating the flow over full-scale aircraft.
  9. Hunsaker, "Forty Years of Aeronautical Research," 251-254; Levine, "U.S. Aeronautical Research Policy," 7-41. The passage in 1926 of the Air Commerce Act, which made the Secretary of Commerce responsible for encouraging and regulating civil aviation, clarified the role of NACA and made possible the focus on aeronautical research.
  10. The great majority of the people who joined the research staff of NACA during the history of the organization, 1915-1958, held degrees in engineering rather than the physical sciences. Thus "research engineer" became the most common formal designation for those working in aeronautical science for NACA.
  11. Gray, Frontiers of Flight, 33-70; Hunsaker, "Forty Years of Aeronautical Research," 254-259. The classic text on subsonic aerodynamics is Richard von Mises, Theory of Flight (2 ed., New York, 1959).
  12. Elsbeth E. Freudenthal, The Aviation Business: Kitty Hawk to Wall Street (New York, 1940), 62-304; John B. Rae, "Financial Problems of the American Aircraft Industry," Business History Review, XXXIX (spring 1965), 99-114.
  13. By 1938 the altitude record set for aircraft, as established by an Italian aviator, had reached beyond 56,000 feet. Eugene M. Emme, Aeronautics and Astronautics: An American Chronology of Science and Technology in the Exploration of Space, 1915—1960 (Washington, 1961), 162.
  14. Hunsaker, "Forty Years of Aeronautical Research," 262.
  15. Levine, "U.S. Aeronautical Research Policy," 74-79; Twenty-third Annual Report of the National Advisory Committee for Aeronautics—1937 (Washington, 1938), 2. The NACA organizational structure, in addition to the 15-member Main Committee, which established the research policies of the agency, and the various field installations, eventually included four technical committees, charged with studying problems in particular areas of aeronautical science and recommending to the Main Committee changes in policy and practice. The membership of the various technical committees, like that of the Main Committee, came from the military, the aircraft industry, and the academic community. Each of the technical committees had subcommittees. In 1957 the technical committees were: Aerodynamics, Power Plants, Aircraft Construction, and Operating Problems. See Forty-third Annual Report of NACA—1957 (Washington, 1957).
  16. Gray, Frontiers of Flight, 19-33; Hunsaker, "Forty Years of Aeronautical Research," 261-262.
  17. Nicholas J. Hoff and Walter G. Vincenti, eds., Aeronautics and Astronautics: Proceedings of the Durand Centennial Conference Held at Stanford University, 5-8 August, 1959 (New York, 1960), 16.
  18. Edgar Buckingham, "Jet Propulsion for Airplanes," in NACA Report No. 159, in Ninth Annual Report of NACA —1923 (Washington, 1924), 75-90.
  19. Hunsaker, "Forty Years of Aeronautical Research," 266-267; Levine, "U.S. Aeronautical Research Policy," 81-89.
  20. See Robert L. Perry, "The Antecedents of the X-1," paper, American Institute of Aeronautics and Astronautics, San Francisco, July 26-28, 1965, 2-17; and Ley, Rockets, Missiles, and Space Travel, 411-413.
  21. Hunsaker, "Forty Years of Aeronautical Research," 267. See also John B. Rae, "Science and Engineering in the History of Aviation," Technology and Culture, III (fall 1961), 391-399. Hunsaker, head of the Department of Aeronautical Engineering at the Massachusetts Institute of Technology and a member of the Main Committee since the 1930s, assumed the chairmanship of NACA in 1941 on Bush’s resignation.
  22. On the role of air power in the Second World War, see Eugene M. Emme, "The Impact of Air Power Upon History," Air University Quarterly Review, II (winter 1948), 3-13; Eugene M. Emme, ed, The Impact of Air Power: National Security and World Politics (Princeton, N.J., 1959), 209-294; and Wesley F. Craven and Janies L. Cate, eds., History of the Army Air Forces in World War II (7 vols., Chicago, 1948-1955).
  23. See C. Fayette Taylor, "Aircraft Propulsion: A Review of the Evolution of Aircraft Powerplants," Report of the Smithsonian Institution for 1961 (Washington, 1962), 245-298.
  24. The best-known of these advisory groups was the so-called von Kármán Committee, established late in 1944 at the direction of Henry H. Arnold, Commanding General of the Army Air Forces, and headed by Theodore von Kármán, of the California Institute of Technology. After surveying wartime achievements in aeronautical science and rocketry, the panel of scientists published its findings in August 1945 and its recommendations in December. While giving full credit to the German accomplishments in rocketry, the von Kármán committee concluded that jet propulsion offered the key to "air supremacy," and that progress toward long-range ballistic missiles should come through the development of air-breathing pilotless aircraft. The philosophy embodied in these 14 reports was to guide Air Force thinking for almost 10 years. See Army Air Forces Scientific Advisory Group, Toward New Horizons: A Report to General of the Army H. H. Arnold (14 vols. [Washington], 1945). For a retrospect of the findings of the committee, see Hugh L. Dryden, "Toward the New Horizons of Tomorrow: First Annual ARS von Kármán Lecture," Astronautics, XII (Jan. 1963), 14-19. Dryden served as deputy scientific director to von Kármán on the committee.
  25. Levine, "U.S. Aeronautical Research Policy," 91-97; Hunsaker, "Forty Years of Aeronautical Research," 267-268.
  26. The unitary plan was designed to provide dispersed NACA-Air Force wind-tunnel facilities characterized by a minimum of overlap and a maximum of variety. Five new supersonic wind tunnels were constructed, one at each of the NACA laboratories and two at a new Air Force installation, the Arnold Engineering Development Center at Tullahoma, Tenn. See Manual for Users of the Unitary Plan Wind Tunnel Facilities (Washington, 1956); and Alan Pope, Wind-Tunnel Testing (2 ed., New York, 1954).
  27. Cite error: Invalid <ref> tag; no text was provided for refs named refc1n27
  28. Cite error: Invalid <ref> tag; no text was provided for refs named refc1n28
  29. Cite error: Invalid <ref> tag; no text was provided for refs named refc1n29
  30. Probably the greatest NACA contribution to the century series (F-100, etc.) was a discovery made in 1951 by Richard T. Whitcomb, an aeronautical engineer working mainly in the recently opened 8-foot, slotted-throat tunnel at the Langley laboratory. Whitcomb collected data on the lengthwise distribution of fuselage and wing volume and suggested an airplane configuration that minimized drag at supersonic speeds. Whitcomb’s findings, known as the “area rule,” indicated that a coke-bottle, or wasp-waisted, shape would significantly increase the speed of jet-propelled airplanes. The importance of the area rule was reflected in the configuration of practically every jet interceptor designed and built for both the Air Force and the Navy in the mid-1950s. See Richard T. Whitcomb, “A Study of the Zero-Lift Drag-Rise Characteristics of Wing-Body Combinations Near the Speed of Sound,” NACA Tech. Report 1273, Forty-Second Annual Report of the NACA–1956 (Washington, 1957), 519-539.
  31. Discussions of the principles of rocketry can be found in many places, but some of the most lucid explanations from the layman’s standpoint are in Ley, Rockets, Missiles, and Space Travel, 60-65; Erik Bergaust and Seabrook Hull, Rocket to the Moon (Princeton, N.J., 1958), 33-43; Ralph S. Cooper, “Rocket Propulsion,” Report of the Smithsonian Institution for 1962, 299-313; and Andrew G. Haley, Rocketry and Space Exploration (Princeton, N.J., 1958), 33-43. See also NASA news release, unnumbered, “Liquid Propellant Rocket Engines,” Jan. 1962. Equally informative as an introduction to rocketry but historically important as a spur to enthusiasts was G. Edward Pendray’s The Coming Age of Rocket Power (New York, 1945), wherein rocket efficiency was pictured as opening “the way to an entire new world of velocities, altitudes, and powers which have hitherto been closed to us; and consequently to a whole new world of human experiences and possibilities” (p. 9).
  32. See A. A. Blagonravov, ed., Collected Works of K. E. Tsiolkovsky, Vol. II: Reactive Flying Machines, NASA TT F-237 (Washington, 1965).
  33. For biographical information on Tsiolkovsky, see A. Kosmodemyansky, Konstantin Tsiolkovsky, His Life and Work, trans. X. Danko (Moscow, 1956); Albert Parry, Russia’s Rockets and Missiles (Garden City, N.Y., 1960), 94-104; Beryl Williams and Samuel Epstein, The Rocket Pioneers on the Road to Space (New York, 1955), 52-69; Heinz Gartmann, The Men Behind the Space Rockets (New York, 1956), 26-35; and K. E. Tsiolkovsky, “An Autobiography,” trans. A. N. Petroff, Astronautics, IV (May 1959), 48-49, 63-64; V. N. Sokolskiy, “The Works of the Russian Scientist-Pioneers of Rocket Technology,” in T. M. Melkumov, ed., Pioneers of Rocket Technology (Moscow, 1964), NASA TT F-9285 (Washington, 1965), 125-162.
  34. Biographical material on Goddard, little known outside of scientific circles until recent years, is accumulating rapidly. A valuable but not definitive biography is Milton Lehman, This High Man: The Life of Robert H. Goddard (New York, 1963). See also E. R. Hagemann, “Goddard and His Early Rockets: 1882-1930,” Journal of the Astronautical Sciences, VII (Summer 1961), 51-59; Eugene M. Emme, “Yesterday’s Dream—Today’s Reality,” Air Power Historian, VII (Oct., 1960), 2 1 6—22 1; G. Edward Pendray, “Pioneer Rocket Development in the United States,” in Emme, The History of Rocket Technology, 19-23; also published in Technology and Culture, IV (Fall 1963), 384-388; Williams and Epstein, Rocket Pioneers, 70-110; Shirley Thomas, Men of Space (6 vols., Philadelphia, 1960-1963), I, 23-46; Gartmann, Men Behind the Space Rockets; and Emme, A History of Space Flight (New York, 1965), 85-87.
  35. Goddard’s 1920 Smithsonian Institution report and a less famous report to the Smithsonian summarizing his findings to 1936 are in Robert H. Goddard, Rockets, Comprising “A Method of Reaching Extreme Altitudes" and "Liquid-Propellant Rocket Development" (New York, 1946). A condensation of Goddard’s notebooks is Esther C. Goddard and G. Edward Pendray, eds., Rocket Development: Liquid-Fuel Rocket Research, 1929–1941 (New York, 1961). The eastern daily newspapers seized on Goddard’s “moon-rocket” reference in his first Smithsonian paper and blew it completely out of proportion. Some journals, having no conception of the mechanics of rocketry, even ridiculed the idea that a rocket could ascend into space, because in a vacuum it would have nothing to “react against.” See, for example, the lead editorial in New York Times, Jan. 13, 1920. The storm of embarrassing publicity doubtless abetted the aversion to notoriety that characterized Goddard throughout his career.
  36. Pendray, “Pioneer Rocket Development in the United States,” 21-23; Pendray, The Coming of Age of Rocket Power, 35-43; Ley, Rockets, Missiles, and Space Travel, 443.
  37. Pendray, “Pioneer Rocket Development in the United States,” 23-24; Pendray, “The First Quarter Century of the American Rocket Society,” Jet Propulsion, XXV (Nov. 1955), 586-593.
  38. Frank J. Malina, “Origins and First Decade of the Jet Propulsion Laboratory,” in Emme, ed., History of Rocket Technology, 52-54.
  39. Ibid., 46—66; Haley; Rocketry and Space Exploration, 97-99; Ley, Rockets, Missiles, and Space Travel, 249-250, 436, 438; Perry, “Antecedents of the X–l,” 20-23.
  40. An exception to the pattern of private sponsorship of rocket societies was the “Group for the Study of Rocket Propulsion Systems,” known as GIRD, established under government auspices in the Soviet Union in 1931. House Committee on Science and Astronautics, 87 Cong., 1 sess. ( 1961 ), House Report No. 67, A Chronology of Missile and Astronautic Events, 3; G. A. Tokaty, “Soviet Rocket Technology,” in Emme, ed., History of Rocket Technology, 275-276; also published in Technology and Culture, IV (Fall 1963), 520-521.
  41. On Oberth see Williams and Epstein, Rocket Pioneers, 111, 143; Gartmann, Men Behind the Space Rockets; Ley, Rockets, Missiles, and Space Travel, 108—130; William Meyer-Cords, “Introduction” to Hermann Oberth, Man into Space: New Projects for Rocket and Space Travel, trans. G. P. H. deFreville (New York, 1957), vii-xiv; Hermann Oberth, “From My Life,” Astronautics, IV (June 1959), 38-39, 100-104; and G. V. E. Thompson, “Oberth—Doyen of Spaceflight Today,” Spaceflight, I (Oct. 1957), 170-171.
  42. Ley, Rockets, Missiles, and Space Travel, 118-162, 197-201; Walter Dornberger, “The German V— 2,” in Emme, ed., History of Rocket Technology, 29-33; also published in Technology and Culture, IV (Fall 1963), 394-395. Williams and Epstein, Rocket Pioneers, 144— 170. Von Braun received a doctorate in physics from the University of Berlin in 1934.
  43. See Walter Dornberger, V-2 (New York, 1954); Dornberger, “The German V-2,” 33-45; Williams and Epstein, Rocket Pioneers, 204-231; Ley, Rockets, Missiles, and Space Travel, 202—231; Dieter K. Huzel, Peenemünde to Canaveral (Englewood Cliffs, N.J., 1962); Leslie G. Simon, German Research in World War II: An Analysis of the Conduct of Research (New York, 1947), 33—35; and Theodore Benecke and A. W. Quick, eds., History of German Guided Missiles (Brunswick, Ger., 1957).
  44. Ley, Rockets, Missiles, and Space Travel, 212-217; Kurt H. Debus, “Evolution of Launch Concepts and Space Flight Operations,” in Ernst Stuhlinger, Frederick I. Ordway III, Jerry C. McCall, and George C. Bucher, eds.. From Peenemünde to Outer Space: Commemorating the Fiftieth Birthday of Wernher von Braun (Huntsville, Ala., 1962), 45. During the powered phase of its flight within the atmosphere the V-2 was stabilized by large aerodynamic fins.
  45. Chronology of Missile and Astronautic Events, 7; Dornberger, “The German V-2,” 32—33. “Vengeance Weapon No. 1”—V–1—was a radio-controlled, subsonic guided missile powered by a pulsejet engine, developed by the German Air Force. Besides the A–4, the accomplishments of the Peenemünde rocket workers included the launching in the early part of 1945 of a winged A–4, called the A–9, which they had designed as the upper stage of a rocket to attack the United States. And by the end of the war Eugen Sanger, already well-known as an Austrian rocket scientist before going to work for the Luftwaffe, and Irene Bredt, a noted German physicist, had written an elaborate report containing a design for an antipodal rocket bomber that would skip in and out of the atmosphere to drop its payload and land halfway around the world. See also Eugen Sanger, Rocket Flight Engineering, NASA TT F-223 (Washington, 1965).
  46. Senate Preparedness Subcommittee of the Committee on Armed Services, 85 Cong., 1 and 2 sess. (1957—58), Inquiry into Satellite and Missile Programs, Hearings, testimony of Wernher von Braun, Part 1, 850; David S. Akens, Historical Origins of the George C. Marshall Space Flight Center (Huntsville, Ala., 1960), 24-29; Tokaty, “Soviet Rocket Technology,” 278-279; James McGovern, Crossbow and Overcast (New York, 1964); Clarence G. Lasby, “German Scientists in America: Their Importation, Exploitation, and Assimilation, 1945-1952,” unpublished Ph. D. dissertation, University of California at Los Angeles, 1962. All together, Paperclip brought nearly 500 aeronautical and rocket scientists, engineers, and technicians to the United States.
  47. Quoted in Tokaty, “Soviet Rocket Technology,” 279.
  48. Inquiry into Satellite and Missile Programs, testimony of von Braun, Part 1, 581; Parry, Russia’s Rockets and Missiles, 118—125.
  49. James R. Baxter, Scientists Against Time (Boston, 1946), 201.
  50. Quoted in Tokaty, "Soviet Rocket Technology", 281.
  51. Ibid., 282-283; Parry, Russia's Rockets and Missiles, 131-133; Frederick I. Ordway III, and Ronald C. Wakeford, International Missile and Spacecraft Guide (New York, 1960), 3-4; Donald J. Ritchie, "Soviet Rocket Propulsion", in Donald P. LeGalley, ed., Ballistic Missile and Space Technology, Vol. II: Propulsion and Auxiliary Power Systems (New York, 1960), 55-85; Chronology of Missile and Astronautic Events, 26; Charles S. Sheldon II, "The Challenge of International Competition", paper, Third American Institute of Aeronautics and Astronautics/NASA Manned Space Flight Meeting, Houston, Nov. 6, 1964.
  52. Among the air-breathing guided missiles (a term that simply meant any pilotless flying craft) designed and developed by the Navy and the Air Force in the first decade after the war were the Gorgon, Plover, Regulus, Cobra, Bomarc, Snark, Matador, and Loon, the last being a Navy version of the German V-1. Of these weapons only the Snark was a genuinely long-range, or intercontinental, missile, and it was subsonic and thus vulnerable to radar-controlled antiaircraft rockets. See Ordway and Wakeford, International Missile and Spacecraft Guide, 3-5, 8-9, 15-16, 20-24, 26, 51.
  53. Quoted, among many other places, in Inquiry into Satellite and Missile Programs, Part I, 283. For a more lengthy argument against early attempts to develop intercontinental ballistic missiles, see Vannaver Bush, Modern Arms and Free Men (New York, 1949).
  54. See Kenneth W. Gatland, Development of the Guided Missile (London, 1954); and Nels A. Parsons, Guided Missiles in War and Peace (Cambridge, Mass., 1956), and Missiles and the Revolution in Warfare (Cambridge, Mass., 1962).
  55. On the postwar V-2 program at White Sands and Cape Canaveral, see U.S. Army Ordnance Corps/General Electric Co., Hermes Project, 1944-1954 (Sept. 25, 1959), 1-4; Ley, Rockets, Missiles, and Space Travel, 254-271; Alkens, Historical Origins of the Marshall Space Flight Center, 28-35; Ernest Krause, "High Altitude Research with V-2 Rockets", Proceedings of the American Philosophical Society, XCII (1947), 430-446; and J. Gordon Vaeth, 200 Miles Up: The Conquest of the Upper Air (2 ed., New York, 1956), 117-134. Unless otherwise indicated, all mileage figures used in this work refer to statute miles. On Thanksgiving Day 1963, several months after Project Mercury officially ended, President Lyndon B. Johnson renamed Cape Canaveral, Cape Kennedy. Since that is beyond the historical context of this study, throughout the rest of this work Cape Canaveral will be used.
  56. Ibid., 178-194; Homer E. Newell, Sounding Rockets (New York, 1959), 54-95; Emme, Aeronautics and Astronautics, 53-54, 58-59, 63, 67, 69-70, 77.
  57. On the Viking see Ley, Rockets, Missiles, and Space Travel, 271-276; Milton Rosen, The Viking Rocket Story (New York, 1955); John P. Hagen, "The Viking and the Vanguard", in Emme, ed., History of Rocket Technology, 123-125; also published in Technology and Culture IV (Fall 1963), 436-437; Vaeth, 200 Miles Up, 195-206; and Newell, Sounding Rockets, 235-242. The first Viking shot, fired in May 1950 from the deck of the Norton Sound in the Pacific, set a new single-stage altitude record, 106.6 miles.
  58. On the Navaho see Ordway and Wakeford, International Missile and Spacecraft Guide, 9-10; and Emme, Aeronautics and Astronautics, 60, 72, 74, 76, 85-86. Besides booster development, the technological heritage from the Navaho program included the airframe for the Hound Dog air-to-surface missile, progress in using titanium for structures, and the guidance system for nuclear-powered submarines.
  59. Alkens, Historical Origins of the Marshall Space Flight Center, 36-37; Wernher von Braun, "The Redstone, Jupiter, and Juno", in Emme, ed., History of Rocket Technology, 108-109; also published in Technology and Culture IV (Fall 1963), 452-455; A. A. Mc-Cool and Keith B. Chandler, "Development Trends in Liquid Propellant Engines", in Stuhlinger, Ordway, McCall, and Bucher, eds., From Peenemünde to Outer Space, 292; John W. Bullard, History of the Redstone Missile System, Hist. Div., Army Missile Command, Oct. 1965, 135-151. The creation of the North Atlantic Treaty Organization in 1949 had provided a clear military need for a battlefield rocket.
  60. Jane's All the World's Aircraft, 1962-1963 (London, 1963), 391-392; von Braun, "The Redstone, Jupiter and Juno," 109-110; McCool and Chandler, "Development Trends in Liquid Propellant Engines," 292; Bullard. "History of the Redstone," 53-93.
  61. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/534
  62. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/534
  63. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/534
  64. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/534
  65. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/534
  66. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/534
  67. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/534
  68. Page:This New Ocean, a history of Project Mercury, Swenson, Grimwood, Alexander (NASA SP-4201).djvu/534