Category Paving the Way for Apollo 11

DEEP-SPACE TRACKING

When the Advanced Research Projects Agency decided in 1958 that a series of probes should be launched towards the Moon, Eberhardt Rechtin led a team at JPL in the development of radio tracking, telemetry and command facilities. When W. H. Pickering created the Telecommunications Division, he made Rechtin its chief. JPL recognised early on that it would need a world-wide network of antennas to maintain contact with deep-space missions. In late 1958 NASA approved the proposal by JPL to develop the Deep Space Instrumentation Facility. To oversee this activity, Abe

Подпись: 6Note that Kuiper, Shoemaker and Urey did not actually originate the experiment.

image31

The Deep Space Instrumentation Facility had large antennas in California, Spain, South Africa and Australia to provide continuous communication with lunar and interplanetary missions.

Silverstein appointed his assistant for Space Flight Operations, Edmond C. Buckley, who had experience of tracking and instrumentation at the rocket range on Wallops Island.7

As the Ranger project geared up in I960, Rechtin was given the additional role of Program Director for the Deep Space Instrumentation Facility. It was decided to build three stations located approximately 120 degrees apart in longitude to provide a continuous line of sight to a spacecraft in deep space. The main station was built near the Goldstone Dry Lake in the Mojave Desert of California, 160 km east of JPL and on the far side of a range of mountains which would shield the antenna from the ‘noise’ of the coastal cities. The other stations were at the Woomera Test Range in Australia and near Johannesburg in South Africa. Later, a fourth station was added near Madrid in Spain.

Because the antennas had to be both large and fully steerable, it was decided to adapt a radio-telescope design. Although the dish was mounted like a telescope, the steering system was designed to hold the antenna pointing precisely at a spacecraft travelling against the background of stars, rather than to maintain sidereal rate. This

In November 1961, James Webb introduced the Office of Tracking and Data Acquisition at NASA headquarters and made Buckley its Director.

was to be done by having the antenna lock onto the spacecraft’s radio transmission and maximise the received signal strength. A 26-metre-diameter dish was required to track the 3-watt transmitter of the Ranger spacecraft. In addition, a system was installed to enable the antenna to simultaneously send ‘uplink’ at one frequency and to receive ‘downlink’ at another. This allowed not only the position of the spacecraft in the sky to be determined, but also both its range and radial velocity along the line of sight. This data would enable the vehicle’s location and motion in space, together known as its state vector, to be monitored continuously in real-time.

Planning for the Space Flight Operations Centre at JPL began in May I960 and was finished in November I960. As it fell within the remit of the Systems Division, Harris Schurmeier appointed Marshall S. Johnson to supervise its construction in Building 125 of the campus. The Space Flight Operations Centre, together with the terrestrial communications network (initially by voice lines and teletype) to link it to the Deep Space Instrumentation Facility, were declared operational on 4 July 1961.

In February 1961 W. H. Pickering ordered a study of future requirements for flight operations, and the recommendation was to construct a new building specifically for this role. In July, NASA gave the go-ahead. This Space Flight Operations Facility entered service in the summer of 1964. Meanwhile, on 24 December 1963 the Deep Space Instrumentation Facility, the terrestrial commu­nications network and the JPL control centre were integrated under the umbrella of the Deep Space Network.8

Note that the Deep Space Network comprised the Deep Space Instrumentation Facility and Space Flight Operations Facility, it did not supersede them.

ON THE SURFACE

The Soviet effort to deliver a capsule to the lunar surface using the rough landing technique finally succeeded with Luna 9. This was launched at 11:42 GMT on 31 January 1966. Its mass was 1,538 kg, including the surface capsule. The midcourse manoeuvre was made at 19:29 on 1 February. As with the Block II Ranger, the inability to deal with a lateral velocity component in the descent limited targets to longitudes of about 64°W and fairly near the equator. In this case, the target was in Oceanus Procellarum, near Hevelius. At an altitude of 8,300 km, with about half an hour to go, the spacecraft aligned its main axis to local vertical. The radar altimeter initiated the retro manoeuvre at 18:44:42 on 3 February, at an altitude of 75 km. At 18:45:30, after slowing by 2.6 km/sec, the engine was cut off when a 5-metre-long probe made contact with the surface, and simultaneously the payload was ejected upward and to the side. The bus hit the ground at 6 m/s and its transmission ceased.

The 250-foot-diameter radio dish at Jodrell Bank in England was the largest fully steerable antenna in the world, and it was monitoring the transmission. When the signal ceased, Bernard Lovell, the head of the facility, wrote it off as another failed landing. But the shock-proof 58-cm-diameter spheroidal capsule rolled to a halt and, some 250 seconds after being released, initiated its own transmission. Four petals opened to right and stabilise the capsule and to expose its contents, which comprised a radiation detector and a line-scan TV camera that pointed upward and viewed the landscape using a nodding mirror that could rotate in azimuth.

Between 01:50 and 03:37 on 4 February a panoramic picture was built up line by line and the data transmitted in real-time. Jodrell Bank recorded the transmission. On a hunch, Bernard Lovell asked the local office of the Daily Express to provide a commercial wire-facsimile machine, and the signal was fed into it. Even before the Soviets announced their probe had transmitted a picture, the ‘scoop’ was published in Britain with the headline: From Luna 9 to Manchester – The Express Catches the Moon. Unfortunately, not knowing how to extract the aspect ratio of the image from the telemetry, they had guessed, and caused the horizontal scale to be compressed by a factor of 2.5, and since it was consistent with the popular expectation of the lunar surface, the resulting jagged landscape seemed ‘right’. The ruggedness was further emphasised by the fact that the Sun was just 7 degrees above the horizon and cast very long, very dark shadows. The surface looked like glassy scoriaceous lava that would be very treacherous for an astronaut to walk on – much like the ‘aa’ lava in Hawaii, so named because a person walking on it tends to cry that sound!

When the official version was issued later using the true aspect ratio, the jagged ‘spikes’ were seen to be just rocks resting on the surface, and the scene was rather less dramatic. The capsule had come to rest oriented 16.5 degrees off vertical. The field of view spanned 11 degrees above and 18 degrees below the perpendicular to the capsule’s axis, with a series of 6,000 vertical lines spanning a full 360 degrees of

image66

A model of the Luna 9 spacecraft showing the spheroidal surface capsule attached to the bus, and (right) the capsule in its deployed configuration. The camera is the cylindrical unit on the axis.

image67

Two sections of a panoramic image transmitted by Luna 9. (Courtesy of Philip J. Stooke, adapted from International Atlas of Lunar Exploration, 2007)

azimuth. As the mirror was only 60 cm off the ground, the perspective was very low, with objects in the foreground appearing larger than they were, and the horizon was very close as a result of the capsule having landed in a shallow 25-metre-diameter crater. There was no sign of the bus.

Gerard Kuiper claimed that there were vesicles in the rocks, which supported his idea that the maria were volcanic. As meteors were particles of dust that penetrated

the Earth’s atmosphere, it seemed only reasonable that the airless Moon would have accumulated a blanket of such material, but this did not seem to be the case. Thomas Gold responded to the evident absence of dust (on this patch of mare, where it could reasonably have been expected to be very thick) by suggesting that the ‘rocks’ were not fragments of lava but fine powder which had adhered to form clods. The surface clearly had sufficient bearing strength to support the capsule’s 100-kg mass – but in the weak lunar gravity its weight was one sixth of this value. To the Apollo planners, this was the most significant result of the mission. Gold argued that the capsule was spreading this load across the four deployed panels, and in time it would sink from sight. The geologists of the Branch of Astrogeology inferred that the surface was (to use Harold Urey’s term) gardened by impacts. The site was on a dark geological unit that Jack McCauley, in making the Lunar Astronautical Chart for this area, had interpreted as a pyroclastic blanket with lava flows. Although there was nothing in the image to suggest pyroclastic, it did indeed look like a lava flow, and judging by the sharpness of the rocks and the absence of dust it was relatively young in terms of lunar history.

A second panorama was taken between 14:00 and 16:54 on 4 February, and this showed that the capsule had increased its tilt to an angle of 22.5 degrees, altering the angle of the horizon. Gold claimed this was evidence of the capsule sinking into the dust. The offset had the benefit of facilitating limited stereoscopic analysis. Before the battery expired on 6 February, further partial pans were made to observe how the illumination changed as the elevation of the Sun in the lunar sky increased, thereby demonstrating the value of repeatedly imaging a scene from a fixed vantage point.

Apollo development

ORGANISATION

On 15 January 1962 the Manned Spacecraft Center reorganised the Apollo Project Office as the Apollo Spacecraft Project Office (ASPO) and appointed Charles W. Frick as Manager, with Robert O. Piland, head of the extinct Apollo Project Office, as Deputy Manager. In recognition of the scale of the project, on 13 February 1963 Piland’s responsibility was narrowed to the LEM and James L. Decker was assigned as Deputy Manager for the CSM.

James Webb again reorganised NASA’s top management on 30 October 1962. In addition to being Director of the Office of Manned Space Flight, Brainerd Holmes became a Deputy Associate Administrator and as such took direct responsibility for the field centres primarily engaged in manned space projects (i. e. the Marshall Space Flight Center, Manned Spacecraft Center and Launch Operations Center) which had previously reported to Robert Seamans. On 20 February 1963, Holmes made George M. Low his Deputy Director for Programs and Joseph F. Shea his Deputy Director for Systems in order to increase their authority over the Directorates of the Office of Manned Space Flight. On 10 May 1963 the Manned Spacecraft Center separated development from operations – as Deputy Director for Development and Programs, James C. Elms was to manage manned space flight projects and plan, organise and direct all administrative and technical support; and as Deputy Director for Mission Requirements and Flight Operations, Walter C. Williams was to manage the writing of mission plans and rules, the training of crews, and the provision of all ground support and mission control facilities. On 9 October 1963 James Webb announced a reorganisation of headquarters to become effective on 1 November. This introduced three Associate Administrators under Robert Seamans. Thus, George Mueller, who on 1 September had replaced Holmes by taking the post of Director of the Office of Manned Space Flight, now also became the Associate Administrator for Manned Space Flight with responsibility for the three field centres most directly involved in manned programs. The Goddard Space Flight Center, JPL and related facilities were assigned to the newly merged Office of Space Sciences and Applications headed by

Associate Administrator Homer Newell. Other facilities were assigned to the Office of Advanced Research and Technology under Associate Administrator Raymond L. Bisplinghoff.

On 22 October 1963 Joseph Shea was reassigned to Houston as Apollo Spacecraft Program Manager. George Low expanded his duties in the Office of Manned Space Flight to include Shea’s post. On 27 August the Manned Space Flight Management Council at headquarters had decided to create a Deputy Associate Administrator for Manned Space Flight Operations so that the Director of the Office of Manned Space Flight could divest himself of this subsidiary role, and on 22 October Mueller drew Walter Williams from Houston for the job.

EARLY TELESCOPIC IMPRESSIONS OF THE MOON

In 1600 William Gilbert, one of Queen Elizabeth I’s physicians and a keen natural philosopher, published the book De Magnete in which he proposed that the planets circled the Sun as a result of some attractive force – and as his great interest was the Earth’s magnetic field he suggested that this force was magnetism. In regard to the Moon, Gilbert shared Leonardo Da Vinci’s view that the bright areas were seas, and after making a sketch of the face of the Moon he assigned names to the dark areas. The drawing was not published until 1651, long after his death.

Hans Lippershey was born in Germany in 1570, but became a citizen of Flemish Zeeland in 1602 and lived in Middleberg, earning his living as a spectacle maker. He is usually credited with discovering in 1608 that using lenses in combination could provide a magnified view of a remote object, but such instruments were apparently developed several times in different places in preceding decades – the telescope was evidently a device whose time had come. On 2 October 1608 Lippershey applied to The Hague for a patent. Several weeks later so did optical instrument maker, Jacob Metius. Both applications were refused. After an account of Lippershey’s telescope was included in a widely circulated diplomatic dispatch, there was a proliferation of telescopes across Europe.

image4

Thomas Harriot began to study the Moon telescopically in the summer of 1609 and in 1611 compiled a map based on his observations.

Thomas Harriot graduated in mathematics from Oxford in England. He tutored Sir Walter Raleigh on navigational issues, in particular the ‘longitude problem’, and on several occasions sailed with him. By 1603 Harriot was wealthy, living in London and pursuing his interest in optics. The appearance of a comet in 1607 prompted his interest in astronomy. In 1608 he obtained from Holland a crude telescope that had a magnification factor of six, and on 26 July 1609 aimed it at the Moon and sketched what he saw. After further observations, in 1611 he compiled a whole-disk map. His work was never published, and it remained unknown until discovered in 1965 by E. Strout of the Institute of the History of Science in the Soviet Union.

Galileo Galilei was born in 1564 in Pisa in Tuscany. He was a mathematician and experimentalist. In 1589 he was made professor of mathematics at the University of Pisa, but three years later took a similar position at the University of Padua. During a visit to Venice in July 1609 he heard of the invention of the telescope via a letter written to one of his friends by a French nobleman. Galileo promptly set out to make one. Whereas Lippershey had used two convex lenses, Galileo combined a convex lens with a concave one to obtain an upright image. It had a magnification factor of ten. On 25 August he displayed it to the Venetian Senate, pointing out that it would enable an inbound ship to be identified several hours earlier than would otherwise be possible – knowledge which would be commercially valuable in a city of merchants. He was rewarded with an increase in salary. In October he went to Florence to show the telescope to his former pupil, Cosimo de Medici, now Grand Duke of Florence. On returning to Padua, Galileo made one with double the magnification factor. On 30 November turned this to the Moon, and again five times over the next 18 nights as the illumination phase changed. As he had a leaning towards painting, his depictions of what he saw were more representational than technical – with the result that few of the features he drew are recognisable. He never drew a full disk to consolidate his observations. After this burst of activity, he seems to have paid little attention to the Moon – but, to be fair, he was busy making discoveries about other celestial bodies. He wrote an account of his observations in the pamphlet Sidereus Nuncius, which he dedicated to de Medici and published on 12 March 1610.

Of the ‘imperfections’ on the Moon he wrote, ‘‘We could perceive that the surface of the Moon is neither smooth nor uniform, nor accurately spherical, […] but that it is uneven, rough, replete with cavities and packed with protruding eminences, in no other wise than the Earth, which is also characterised by mountains and valleys.’’ He was particularly struck by the shadows, which appeared totally black – there was no detail evident within them. He used the shadows cast by mountains to estimate their heights.[1]

But Galileo’s most significant discovery was that Venus displays phases similar to

image5

Representations of the Moon at two illumination phases made by Galileo Galilei in late 1609 and published in 1610.

the Moon, which was proof that this planet orbits the Sun, not Earth. Although the Church was not overly concerned about imperfections on the Moon, it was firmly of the belief that Aristotle was correct in saying that Earth was located at the centre of a system of crystal spheres. However, the phases of Venus meant that Copernicus was correct, which was a serious matter. After being hounded by the Roman Inquisition, in 1633 Galileo was obliged to publicly “curse and detest” the false opinion that the Sun held the central position. He was then placed under house arrest, and that is how he lived until his death in 1642.

Although Galileo may not have been the first to aim a telescope at the heavens, he was the first to publish, and the rapid distribution of his pamphlet prompted a great many people to obtain instruments to look for themselves.

Francesco Fontana, a Neapolitan lawyer, began to observe the Moon in 1630, and in 1646 published Novae Coelestium, Terrestriumque Rerum Observationes, featuring wood-cut engravings of two of his drawings made at different illumination phases.

Although Kepler knew that Galileo’s findings confirmed the Sun to be centrally located, he was wary of saying so explicitly. Since the Moon was evidently a world in its own right, he wrote an allegorical fantasy, Somnium, in which he related how ‘demons’ transferred his hero character to the Moon at the time of a lunar eclipse by sending him down the Earth’s shadow. The fact that the explanatory footnotes were longer than the text of the story established it to be a technical treatise disguised as a work of fiction. Even so, it was not published until 1634, four years after his death.

Jeremiah Horrocks in England observed in 1637 the dark limb of the Moon occult the stars of the Pleiades cluster, one by one. If the Moon possessed an atmosphere, the starlight would have flickered and faded as it was attenuated and refracted by the gas, but in each case the star’s disappearance was instantaneous.

In 1638 John Wilkins, an English clergyman, published Discovery of a World in

the Moone; Or a Discourse tending to prove that ’tis probable there may be another Habitable World in that Planet. He took a serious look at how a voyage to the Moon might be attempted utilising some form of ‘engine’. Wilkins was so enthusiastic that after he helped to establish the Royal Society of London in 1660 he had this petition the government to undertake such a venture with the objective of claiming the Moon for the British Empire!

The Apollo commitment

THE ROLE OF MAN IN SPACE

In 1958 NASA was assigned the task of exploring space for scientific purposes, but no immediate objectives were specified. The National Aeronautics and Space Act left the agency to set its own goals.

Accordingly, within days Keith Glennan established the Space Task Group at the Langley Research Center to manage Project Mercury, which was to launch a man into orbit as soon as possible. This was not exactly what President Eisenhower had had in mind, but he saw it as a one-off venture. At a packed press conference on 9 April 1959, Glennan introduced the seven military test pilots who had been chosen to be astronauts.[16]

On 25-26 May 1959 Harry J. Goett of the Ames Research Center chaired the first meeting of the Research Steering Committee on Manned Space Flight, which was to consider possible man-in-space objectives for the coming decade. These included:

• launching and operating a small orbital laboratory

• assembling a large permanent space station

• flying circumlunar and lunar orbital missions

• making a lunar landing.

George M. Low, who represented headquarters, was firmly of the view that only a manned lunar landing provided a reasonable ultimate objective – it was an ‘end’, not just an intermediate step along a path. Whilst less demanding, the alternatives were not as definitive – either a man landed on the Moon or he did not; there was no way

to redefine it as something simpler and assert this to be equivalent.[17] It was therefore decided to set the long-range goal of achieving a manned lunar landing in the 1970s. This would gave a framework in which to define a series of intermediate objectives for the 1960s.

On 5 July 1960 the House Committee on Science and Astronautics said, “NASA’s 10-year program is a good program as far as it goes, but it does not go far enough.’’ In other words, it urged NASA to accelerate its long-range plan. In the committee’s view, “A high priority program should be undertaken to place a manned expedition on the Moon in this decade. A firm plan with this goal in view should be drawn up and submitted to the Congress by NASA.’’ But it warned that this plan, “should be completely integrated with other goals, to minimise total costs. The modular concept deserves close study. Particular attention should be paid immediately to long lead­time phases of such a program.’’

Eisenhower responded by asking his science advisor, James R. Killian, whether a manned lunar landing represented a scientific venture which could be justified in the same manner as launching a satellite for the International Geophysical Year. Killian convened a meeting of scientists, and their report, written by Donald F. Hornig of Princeton University, was dismissive: “At the present time, man-in-space cannot be justified on purely scientific grounds.’’ The rationale for sending men to the Moon seemed to be “emotional compulsion and national aspirations’’. Hence Eisenhower refused funding for manned space flight beyond Project Mercury, and in particular the proposal for a three-man Apollo spacecraft. He had no problem with the agency using the new Saturn booster to launch heavy satellites, but he withdrew funding for the upper stages intended to enable this to launch a manned spacecraft. Nevertheless, on 12-13 September the Space Task Group held a briefing for potential bidders to develop the Apollo spacecraft with the Moon as the ultimate objective, and released the formal request for proposals. On 17 October, Low told Abe Silverstein, Director of the Office of Space Flight Programs, that he was going to set up a committee to study the circumlunar objective in greater detail, to ensure that the Apollo spacecraft would be capable of supporting a landing mission.[18] On 25 October NASA issued contracts to three companies to provide feasibility studies for the Apollo spacecraft.

The national election of November 1960 was won by John F. Kennedy. He was inaugurated on 20 January 1961. In giving his final budget speech prior to leaving office, Eisenhower said on 18 January 1961 that Congress would have to determine “whether there are any vital scientific reasons for extending manned space flight beyond Mercury’’. In a campaign statement, Kennedy had said: “We’re in a strategic space race with the Russians, and we are losing. If a man orbits Earth this year, his name will be Ivan. If the Soviets control space they can control the Earth, as in past centuries the nation that controlled the seas has dominated the continents. We

cannot afford to run second in this vital race. To insure peace and freedom we must be first. Space is our great New Frontier.” The contrast with Eisenhower’s view was stark. Kennedy also had an appreciation of national prestige, which in the Cold War meant a comparison with the achievements of the Soviet Union. The issue of prestige had been dismissed by Eisenhower.

During the transitional period, Kennedy assigned a number of task forces to draw up policy recommendations. The Committee on Space was chaired by Jerome B. Wiesner, who served on the President’s Science Advisory Committee during James Killian’s chairmanship and was to become Kennedy’s Special Assistant for Science and Technology. The Committee on Space in turn set up the Panel on Man-in-Space, composed largely of scientists, and its report on 18 November I960 criticised the program envisaged by NASA.4 Although it agreed the need for large launch vehicles and urged an emphasis on space science and applications, it criticised “the popular belief that man in space is the most important aim of our non-military space effort’’. Wiesner recommended that Project Mercury be ended as soon as it had achieved its objective of placing a man in orbit, and that there should be no follow-on. However, Kennedy had made Vice President Lyndon B. Johnson chairman of the National Aeronautics and Space Council, and Johnson was in favour of expanding the space program.

And when the Space Science Board of the National Academy of Sciences issued a position paper on Man’s Role in the National Space Program on 27 March 1961 it said, “scientific exploration of the Moon and planets should be clearly stated as the ultimate objective of the US space program for the foreseeable future. This objective should be promptly adopted as the official goal of the United States space program and clearly announced, discussed and supported.’’ It also advised that whilst it was “not now possible to decide whether man will be able to accompany expeditions to the Moon and planets’’, NASA should proceed with its planning “on the premise that man will be included’’. Taking the broader view, it said that such exploration would be “potentially the greatest inspirational venture of the century and one in which the entire world can share; inherent here are great and fundamental philosophical and spiritual values which find a response in man’s questing spirit’’. Clearly this national scientific body, established to advise NASA on policy, was taking a much broader view than the sky scientists involved in space research at that time.

IN LUNAR ORBIT

Having achieved the first delivery of a scientific capsule to the lunar surface, the Soviets moved on to attempt to be first to put a satellite into orbit around the Moon. Luna 10 was launched at 10:47 GMT on 31 March 1966. After cruising in parking orbit, it set off for the Moon. It was the same type of bus as Luna 9, but was ferrying an instrument capsule instead of the landing capsule. The midcourse manoeuvre on 1 April refined the trajectory to aim for the point in space at which the 850-m/s orbit insertion burn would be made. When 8,000 km out, it oriented itself for braking. The burn was initiated at 18:44 on 3 April, and slowed the spacecraft sufficiently for it to enter a 350 x 1,017-km orbit with a period of 178 minutes, in a plane inclined at 72 degrees to the lunar equator. The fact that the change in velocity to enter orbit was considerably less than that to land meant that propellant could be traded in favour of an increase in the payload to 245 kg. Shortly after entering orbit the bus released the 1.5-metre-long capsule containing a micrometeoroid detector, radiation detectors, an infrared sensor to measure the heat flux from the Moon, a gamma – ray spectrometer to detect radioactive isotopes, and a magnetometer to follow up the measurements by the early flyby probes. The mission ended when the battery expired on 30 May, after 56 days during which the capsule made 460 revolutions.

image68

In lunar orbit

141

The gamma-ray spectrometer was similar to that of the Ranger Block II, but more useful by virtue of being placed into orbit to survey a wide area. The instrument was a scintillation spectrometer with a resolution of 32 channels within the energy range 0.3-3.0 MeV. The surface resolution was rudimentary. The data was consistent with the proposition that the maria were basaltic, but was inconclusive. About the only positive conclusion was there were no large surface exposures of acidic rock such as granite. The question for the ‘hot Moon’ hypothesis advocated by Gerard Kuiper, was why the process which produced ‘continental’ material on Earth had seemingly not done so on the Moon. The mystery was the composition of the highland material. If a global magnetic field existed, then it was weaker than 1/100,000th that of Earth. Radio occultations on crossing the limb showed no hint of the Moon having even a tenuous envelope of gas. Intriguingly, radio tracking revealed the gravitational field to be uneven.

PLANS AND SCHEDULES

In view of the reason for his predecessor’s resignation, George Mueller ordered a review of Apollo, and this confirmed the project to be in trouble. On 29 October 1963 Mueller informed the Manned Space Flight Management Council that the only way to recover time would be to reduce the number of development flights. The plan drawn up by the Marshall Space Flight Center in March 1962 envisaged a series of launches of the Saturn V in which the stages were tested in sequence – with only the first stage being ‘live’ on the first test scheduled for late 1965. The aim was to ‘man rate’ this vehicle by the summer of 1967, then use it to launch at least six manned missions in Earth and lunar orbit prior to attempting a lunar landing in late 1968 or early 1969. Mueller proposed to reduce this research and development phase by ‘all up’ testing in which each launch would use only ‘live’ stages, modules, systems and spacecraft. Wernher von Braun and Robert Gilruth objected, but Mueller had the support of James Webb.

In addition, a recent study by Bellcomm had recommended reassigning the early tests of the Apollo Block I spacecraft from the Saturn I to the Saturn IB, and so on 30 October Mueller cancelled the four manned test flights with the Saturn I that had been set for 1965. The development of the Saturn IB for manned missions would be accelerated and the ‘all up’ testing strategy employed in this case too. After coming to terms with this, Gilruth asked von Braun whether the Saturn IB could lift both the CSM and LEM, and was advised that it would be feasible only if their weights were controlled. At the White Sands Missile Range in New Mexico on 7 November the Apollo launch escape system successfully performed its first ‘pad abort’ test. On 18 November 1963 Mueller directed that if the LEM was not ready in time, the early Saturn IB flights would fly without it. But it must be phased into the test program as quickly as possible. Furthermore, Mueller directed that two successful development flights for each of the Saturn IB and Saturn V would serve to ‘man rate’ them. The schedule that he issued on 31 December 1963 listed the first Saturn IB test in early 1966 and the first manned mission later that year. The first Saturn V test was to be in the first quarter of 1967, with the first manned flight (hopefully on the third launch) later that year. Mueller then established the Apollo Program Office with himself as Director, and hired Samuel C. Phillips, who had managed the development of the Air Force’s Minuteman missile, as Deputy Director.

On 17 September 1962 NASA had announced the nine men of its second intake of astronauts.1 At the same time, Deke Slayton was appointed Coordinator of Astronaut Activities, reporting to Robert Gilruth.[41] [42] In addition to the administrative tasks of the Astronaut Office, which Slayton managed in the manner of a military unit, he was responsible for making flight crew assignments. On 18 October 1963 the fourteen men of the third astronaut group were announced.[43] By now James Elms was Deputy Director of the Manned Spacecraft Center, and on 5 November 1963 Gilruth inserted Assistant Directors under Elms in order to strengthen the local management of flight operations: Chris Kraft was redesignated as Assistant Director for Flight Operations, Deke Slayton as Assistant Director for Flight Crew Operations and Maxime Faget as Assistant Director for Engineering and Development. In addition, Merritt Preston was assigned to manage Manned Spacecraft Center operations in Florida. However, on 17 January 1964 Elms resigned, and two days later George Low was reassigned from headquarters to replace him.

President Kennedy flew to Cape Canaveral on 16 November 1963 to inspect the ‘moonport’ which NASA was beginning to construct on nearby Merritt Island. He was shown models to illustrate the enormous size of the Saturn V. On 22 November he was assassinated in Dallas, Texas, and later that day Lyndon Johnson was sworn in as his successor. In a TV address on 28 November Johnson directed that Cape Canaveral be renamed Cape Kennedy, and the next day he signed an executive order in which the Launch Operations Center was renamed the John F. Kennedy Space Center.[44]

On 15 January 1964 the Manned Spacecraft Center proposed to Apollo Spacecraft Program Manager Joseph Shea that two of the Saturn IB ‘all up’

Group 1, seated (left to right): Captain Leroy Gordon Cooper Jr, Captain Virgil Ivan ‘Gus’ Grissom, Lieutenant Malcolm Scott Carpenter, Lieutenant Commander Walter Marty Schirra Jr, Lieutenant Colonel John Herschel Glenn Jr, Lieutenant Commander Alan Bartlett Shepard Jr and Captain Donald Kent ‘Deke’ Slayton. Group 2, standing (left to right): Captain Edward Higgins White II, Captain James Alton McDivitt, Lieutenant Commander John Watts Young, Elliot McKay See Jr, Lieutenant Charles ‘Pete’ Conrad Jr, Major Frank Frederick Borman II, Neil Alden Armstrong, Captain Thomas Patten Stafford and Lieutenant Commander James Arthur Lovell Jr.

Group 3, seated (left to right): Major Edwin Eugene ‘Buzz’ Aldrin Jr, Captain William Alison Anders, Captain Charles Arthur Bassett II, Lieutenant Alan LaVern Bean, Lieutenant Eugene Andrew Cernan and Lieutenant Roger Bruce Chaffee; standing (left to right): Captain Michael Collins, Ronnie Walter Cunningham, Captain Donn Fulton Eisele, Captain Theodore Cordy Freeman, Lieutenant Commander Richard Francis Gordon Jr, Russell Louis ‘Rusty’ Schweickart, Captain David Randolph Scott and Captain Clifton Curtis Williams.

During a visit to Cape Canaveral on 16 November 1963 John F. Kennedy is briefed by George E. Mueller on the ‘mobile launcher’ concept for Apollo. To Kennedy’s right are (in turn) James E. Webb, Robert C. Seamans, Kurt H. Debus and George M. Low. To his left are Hugh L. Dryden, Wernher von Braun, General Leighton I. Davis and Florida Senator George A. Smathers.

development flights be used to test the heat shield of the Apollo command module, because this would enable the early tests of the Saturn V to be classified as ‘demonstration’ rather than ‘development’ for the spacecraft.[45] On 7 February Grumman was directed to provide two LEM test articles (LTA) and eleven flightworthy LEMs, the first three of which were to be capable of either manned or unmanned operation. On 23 March George Mueller ordered that if the first two unmanned CSM test flights were successful, the next mission would be a long- duration manned flight, after which there would be two tests of the LEM, the first

Table 15.1: Outline schedule for Apollo drawn up in November 1964

Mission

Payload

Launch Date

AS-201

CSM-009 (unmanned)

1965

AS-202

CSM-011 (unmanned)

1966

AS-203

No spacecraft (S-IVB development flight)

Jul 1966

AS-204

CSM-012 (manned)

Oct 1966

AS-205

CSM-014 (manned)

Jan 1967

AS-206

LEM-1 (unmanned)

Apr 1967

AS-207

CSM-101 (manned) and LEM-2

Jul 1967

one unmanned and the second together with a manned CSM – so long as the Saturn IB proved capable of lifting both vehicles together. In November 1964 Joseph Shea, George Mueller and Sam Phillips drew up an outline schedule for testing Apollo hardware in advance of the introduction of the Saturn V, but it remained uncertain whether the weights of the two spacecraft were sufficiently constrained for them to be lifted together by a Saturn IB for the joint mission. On 16 December Shea directed that the Block I manned missions must use low orbits from which the spacecraft could use its reaction control system thrusters to de-orbit itself in the event of the failure of the service propulsion system; and in the event of these too failing, the orbit must decay naturally and result in re-entry within an acceptable duration.6

On 31 August 1964 Lead Flight Director Chris Kraft appointed John D. Hodge, Eugene F. Kranz and Glynn S. Lunney to alternate in round-the-clock flight operations. On 24 December Everett E. Christensen was made Director of Mission Operations, a position which effectively superseded Deputy Associate Administrator for Manned Space Flight Operations – vacant since the resignation of Walter Williams in April. At the same time, two posts of Mission Director were also created, with the intention that the appointees would run alternate missions. In addition, activities at the Cape were consolidated, with Kurt H. Debus being made Director of Launch Operations and Merritt Preston, who had been managing the Manned Spacecraft Center’s activities at the Cape, becoming his Deputy.

Joseph Shea, Chris Kraft and Deke Slayton were briefed on 18 January 1965 by the Mission Planning and Analysis Division of the Manned Spacecraft Center about the Saturn IB and early Saturn V flights. On 21 January, in response to a question by Sam Phillips, Shea said the current estimate was that the Saturn IB would be able to insert 35,500 pounds into a circular orbit at 105 nautical miles. This, however, was less than the combined ‘control weights’ of the CSM and LEM by 870 pounds, and both vehicles were currently above their control weights. Shea argued that in view of the difficulty in constraining the weights, the best solution would be to find a way of increasing the launcher’s capacity by 1,000 pounds. In fact, the Saturn IB had a

This precaution was reputedly a headquarters response to the situation depicted by Martin Caidin in his recent novel Marooned.

‘control payload’ which was the specified minimum mass that it was to be capable of placing into the reference orbit, and a ‘design goal’ which exceeded this. On 23 February Phillips told Shea that the Marshall Space Flight Center would endeavour to increase the payload by 1,000 pounds. The development version of the cluster of eight H-1 engines had yielded 1.3 million pounds of thrust, but the fifth flight of the Saturn I had introduced an upgraded cluster that finally achieved its specification of 1.5 million pounds of thrust. In August 1963 Rocketdyne had proposed an upgrade for 1.6 million pounds of thrust, and on 8 November of that year NASA had ordered this be done. By 23 April 1965 the improved engine had completed its qualification testing. On 12 May Huntsville reported that it would be possible to uprate the engine by an additional 5,000 pounds of thrust, to raise the total to 1.64 million pounds. But the rocket engineers were fighting a losing battle, as by then both spacecraft had put on even more weight.

On 13 January 1965 Shea had established the Configuration Control Board, with himself in the chair. This was to rule on all proposals for engineering changes to the spacecraft. On 10 May he faced a dilemma: the ‘all up’ testing regime required that all spacecraft incorporate a full set of subsystems, but it had been proposed that the landing radars be omitted from LEM-1 and LEM-2 on the basis that a radar would serve no function on an Earth orbital mission. Omitting the radar on these early test flights would save money and assist in the effort to trim the weight of the vehicle at this critical juncture, but doing so would establish the precedent for a series of one – of-a-kind spacecraft, each tailored to achieving specific development objectives and with none demonstrating all of the systems in conjunction. Furthermore, by relieving the pressure on the effort to trim weight in the short term, such compromises might jeopardise it in the long term. On 27 May the Manned Spacecraft Center reaffirmed that LEM-1 must test the radar. But on 25 June ASPO Assistant Manager Harry L. Reynolds warned Owen E. Maynard, Chief of the Systems Engineering Division at the Manned Spacecraft Center, that it was ‘‘becoming increasingly clear that we are going to have a difficult job keeping the LEM weight below the control weight’’. On 6 July Grumman requested to be allowed to deliver the early LEMs without some subsystems installed, but Shea insisted they must all leave the factory in a fully functional condition. At that time, LEM-1 was to be delivered to the Kennedy Space Center in November 1966, with the next five vehicles following in 1967, but it was becoming increasingly evident that this schedule would be difficult to achieve. On 13 September 1965 Shea established the Weight Control Board to enable subsystem managers to meet on a weekly basis and report progress in controlling the weights of the two spacecraft, and when appropriate to create ad hoc task forces to chase up specific issues and report back.

Meanwhile, on 17 February 1965 Shea clarified for North American Aviation the Block I schedule. CSM-009 and CSM-011 were to be configured for unmanned use and fly as AS-201 and AS-202 to test the heat shield. CSM-012 and CSM-014 were to be delivered for manned missions, but be capable of being adapted at the Cape for unmanned flight. The decision for CSM-012 would be made 6 months ahead of the scheduled launch date for AS-204, and if flown unmanned this would be done either to gain additional data on the spacecraft’s characteristics or to provide more time for

the Marshall Space Flight Center to prepare AS-203 to obtain data on the behaviour of the S-IVB stage in space. North American Aviation was told that CSM-017 and CSM-020, assigned to the early tests of the Saturn V, need not be capable of manned use. The first manned Block II would be CSM-101, which was to fly in conjunction with LEM-2. On 22 March Glynn Lunney, Chief of the Flight Dynamics Branch of the Flight Control Division in Houston, was appointed Assistant Flight Director for AS-201 and AS-202. On 25 June Carroll H. Bolender was made Deputy Director of Mission Operations at the Office of Manned Space Flight, and his first task was to plan these two preliminary missions.

On 10 August 1965 ASPO named LEM-1 to AS-206, LEM-2 to AS-207, LEM-3 to AS-503, LEM-4 to AS-504, LEM-5 to AS-505 and LEM-6 to AS-506. Of the six test articles, LTA-1 was kept by Grumman at Bethpage to resolve issues during the initial fabrication, assembly and checkout procedures, LTA-2 went to the Marshall Space Flight Center for launch vibration tests, LTA-3 and LTA-5 were to be used to assess the structural effects of engine firing, LTA-8 went to the Manned Spacecraft Center for thermal-vacuum environmental testing, and LTA-10 went to the North American Aviation factory in Tulsa, Oklahoma, for fit-checks with the SLA, which was being manufactured there. To cut costs, in July Grumman had been directed to delete LTA – 4 (intended for vibration tests), the ascent stage of LTA-5 and the two flight test articles and instead to refurbish two of the test articles for flight once their ground testing role was complete. The company said it would refurbish LTA-10 and LTA-2 in case they were needed for the first two Saturn V test flights. The first three LEMs were to incorporate development flight instrumentation so as to record the dynamic environment at launch. A key requirement was that the differences between LEM-3 and LEM-4 be minimised and that all subsequent production vehicles be identical.

On 21 October 1965 Sam Phillips slipped AS-201 to January 1966 and AS-202 to June 1966 to accommodate the revised delivery dates for CSM-009 and CSM-011, but otherwise preserved the outline schedule which had been drawn up in November 1964. On 2 December 1965 Hugh Dryden died of cancer.[46] Robert Seamans replaced him as Deputy Administrator on 21 December. He retained the duties of Associate Administrator until Homer Newell gained this post in August 1967, and was in turn superseded as Associate Administrator for Space Sciences and Applications by John E. Naugle.

Meanwhile, an operational step toward the chosen Apollo ‘mission mode’ was achieved when Gemini 6 rendezvoused with Gemini 7 on 15 December 1965. The straightforward manner in which this was done raised the prospect of undertaking the manned test of the LEM without reducing the weights of the CSM and LEM to enable the Saturn IB to lift them both together. On 28 January 1966 Sam Phillips asked ASPO to assess the impact, including the effects on ground support equipment and mission control, of a dual AS-207/208 mission as early as the scheduled date for

AS-207, which was the Saturn IB that was nominally to have sent them into orbit together. The idea was for near-simultaneous launches of AS-207 with CSM-101 and AS-208 with LEM-2 to facilitate a rendezvous and docking, at which point the mission would unfold as originally planned. On 2 February John P. Mayer, Chief of the Mission Planning and Analysis Division at the Manned Spacecraft Center, informed Chris Kraft, Assistant Director for Flight Operations, that the main constraint would be programming the Real-Time Computer Complex in Houston to plan and support such a mission – in which case the decision on whether it was to be attempted must be taken very soon. Mayer also urged that if the IBM staff who worked on the Gemini 6/7 rendezvous could be spared, they should be reassigned to help to plan the new dual mission. On 4 February John Hodge, Chief of the Flight Control Division, noted that some of the operational issues associated with near­simultaneous launches would be obviated if the interval were extended. On 24 February Mayer’s assistant, Howard W. Tindall, recommended that the CSM be launched first and the LEM follow it either 24 hours later or at a recurring daily window. On 1 March Joseph Shea endorsed the concept. On 8 March Sam Phillips directed the Manned Spacecraft Center, Marshall Space Flight Center and Kennedy Space Center to endeavour to launch the dual mission a month later than intended for AS-207 on the previous schedule.

MAPPING THE MOON

Jacob Floris van Langren founded a business in Amsterdam in 1586 which made globes, and as Dutch explorers reported discoveries he could barely keep up with the demand for updates. In 1627 his grandson, Michel van Langren, observed the Moon and made a sketch. After moving to Madrid as Court Astronomer to King Felipe IV of Spain in 1630, the grandson convinced the King that tables listing the sunrise and sunset times of specific lunar features would enable the time at the observing site to be determined, which would in turn solve the ‘longitude problem’. The prerequisite was a map of the Moon. In 1643, having made 30 sketches, van Langren realised he had competitors, so in 1645 he issued a whole-disk map 34 cm in diameter on which he named 325 features after prominent philosophers, mathematicians, astronomers, explorers, religious figures and (recognising his sponsor) members of the Spanish royal family. However, at that time the revolt of Protestantism which would later be called the Thirty Years War was well underway, and a nomenclature drawn from Catholic Europe was sure to be contentious.[2]

In 1637 Pierre Gassendi, a mathematician in Paris, also came to the conclusion that it should be possible to use observations of the Moon to determine the time and thereby resolve the ‘longitude problem’. After he had made some drawings, he heard that Johann Hevelius, whom he had once met, was starting to make a map, and upon seeing the quality of the younger man’s sketches Gassendi stopped and handed over his own work. A city councillor in Danzig in Poland, Hevelius built an observatory on the roof of his house and installed a telescope with a 5-cm-diameter lens, a focal length of 3.6 metres and a magnification of 50 – in fact, one of the best telescopes of the time. In 1647 he published Selenographica sive Lunae Descriptio, with fine drawings and a consolidated map 30 cm in diameter.[3] He named 275 features after terrestrial landforms, including oceans, seas, bays and lakes – although he realised there were no bodies of open water. Like Galileo, Hevelius estimated the heights of the lunar peaks by their shadows, but much more accurately. Being a Protestant, his nomenclature had little in common with that of van Langren. In fact, Hevelius had presumed himself to be the first to name features, and said the task was arduous. He

image6Ш7^ f

Подпись: . ЯПС^ммип Mtfr uhiufi»tвШ ь. ftti ttviat fiut W.VTW й^п,|ні тій<<_» ~~,1'“' “—[упчУіІігг/і - УДЧІХГ nwr/, I* Ьш*і« at Подпись:

MAPPING THE MOON
Подпись: rc"««.'4vcv
Подпись: .tn'tllV.

ШІІІ »Г«Ш л*1" V«…

‘ Піт.

Д-U^li iM ЛкЖсЛіШІ »Wfca«£e ■

ши: ГОфАМ WUJA ГЛд»>Т1П TAMES SVXASO
мшмліичшоуіампіА.

MICHAEL TLORENTIVe. LANGRENVS

The map of the Moon published by Michel van Langren in 1645 was the first to assign names to features.

image7

A map of the Moon published by Johann Hevelius in 1647.

honoured astronomers and scientists, including Gassendi, but not himself. However, only a dozen of his names have survived.

As a Jesuit professor of astronomy and theology at the University of Bologna, Giovanni Battista Riccioli believed implicitly in the Aristotlean system as written of by Ptolemy. In an effort to counter the growing belief that Earth travels around the Sun, he set out to write an authoritative account of astronomy. But while developing his argument he came to suspect he was wrong! He could never admit this publicly, however. In 1651 he published Almagestum Novum, with a whole-disk map of the Moon that was 28 cm in diameter and was based on observations made by his pupil, Francesco Maria Grimaldi. Although the map was little better than that by Hevelius, its historical significance was the nomenclature. This retained oceans, seas and bays for the dark areas, but renamed them for states of mind: e. g. Oceanus Procellarum and Mare Tranquillitatis. Craters were named after astronomers and philosophers, including Riccioli and Grimaldi. The despised Copernicus was assigned a crater in Oceanus Procellarum – the Ocean of Storms. To Helvelius’s frustration, soon copies of his map were in circulation relabelled with Riccioli’s nomenclature! Nearly all of the 200 names introduced by Riccioli and Grimaldi are still in use today.

Giovanni Domenico Cassini was born in 1625 in the Republic of Genoa. After a Jesuit education he was hired by the Marquis Cornelio Malvasia in Bologna, who

image8

A map of the Moon published by Giovanni Battista Riccioli in 1651.

derived ephemerides for astrological purposes. Utilising the excellent instruments of his employer’s observatory, Cassini made observations of exceptional precision and quality, and in 1650 became professor of astronomy at the University of Bologna. In 1666 plans were initiated to establish a national observatory in Paris, and in 1669 Cassini, now with several significant discoveries to his name, was invited by King Louis XIV to become its first director; he accepted and promptly moved to France to oversee the construction of the observatory, which was finished in 1671. In 1679 he published a map of the Moon which, at 52 cm in diameter, was much larger than its predecessors. Although very accurate, so few copies were made that it did not gain the attention which it warranted.

Meanwhile, Isaac Newton at the University of Cambridge in England had made a study of gravity and, contrary to the accepted wisdom that it remained constant

image9

A map of the Moon published by Giovanni Domenico Cassini in 1679.

with distance, realised that its strength declined with the inverse square of distance. His book Philosophiae Naturalis Principia Mathematica, published in 1687, provided a basis for the laws of planetary motion that Kepler had derived empirically.

The Moon was essentially ignored for a century, then Tobius Mayer in Germany became interested in its use in relation to the ‘longitude problem’. In 1751 he gained the chair of economics and mathematics at the University of Gottingen, where, a few years later, he became superintendent of the observatory. As a skilled draughtsman, he utilised a micrometer to measure the geographical positions of the lunar features. His map was published posthumously in 1775, and although only 20 cm in diameter it was the first to include lines of latitude and longitude. It superseded the map by Hevelius (re-annotated with the nomenclature of Riccioli) which had been standard for almost 150 years, and would itself not be surpassed for half a century.

image10

A map of the Moon by Tobius Mayer that was published posthumously in 1775.

William Herschel, who discovered the planet Uranus in 1781 and became the first president of the Royal Astronomical Society, did not devote much attention to the Moon. However, he believed it to possess an atmosphere (even though the way stars were occulted contradicted this) and active volcanoes. Although by now the idea of open water had been abandoned, Herschel was “absolutely certain” the Moon was inhabited.

Johann Hieronymus Schroter was born in 1745 in Erfurt in Germany. In 1767 he graduated in law from the University of Gottingen. In 1781 he moved to Lilienthal, near Bremen, to become chief magistrate. Having been inspired by Mayer’s map, he built an observatory alongside his house and installed a series of ever more powerful telescopes. Over a period of 30 years he made hundreds of detailed drawings of the Moon, recording individual features under different angles of illumination. In this respect, he founded modern selenography. He paid particular attention to craters and

rilles – both of which terms he introduced. His measurements of the heights of lunar peaks were better than those of his predecessors. Although an accurate observer, he was not a skilled draughtsman and utilised a ‘schematic’ style. In 1791 he published Selenotopographische Fragmente zur genauern Kenntniss der Mondflache in two volumes containing a total of 75 engravings. He did not consolidate his observations into a full – disk map, but did include an enlarged version of Mayer’s map. In 1813 the invading Napoleonic army ransacked and destroyed his observatory, and most of his unpublished work was lost. Schroter inferred the Moon to have an atmosphere, but estimated its pressure to be less than that of the best vacuum pump available at that time. Nevertheless, like Herschel, he believed the Moon to be inhabited.

Wilhelm G. Lohrmann was a surveyor in Dresden, and a skilled draughtsman. He had developed a keen interest in astronomy as a boy, and in 1821, aged 25, began to study the Moon. He set out to make a map in 25 sections on a scale at which the disk would be 95 cm in diameter. By 1824 he had released the first four sections, and by 1836 had the drawings for the remaining sections, but then his eyesight failed and he was unable to finish the editing. Nevertheless, in 1838 he published a full-disk map at 40 per cent scale.

Johann von Madler was born in 1794 and became a teacher in Berlin. One of his students, the wealthy banker Wilhelm Beer, was only a few years younger and they became friends. In 1829 Beer built an observatory at his house and bought a 95-mm – diameter refracting telescope produced by Joseph von Fraunhofer’s firm. Beer hired Madler as observer, and they measured almost 1,000 features to trigonometrically survey the Moon. Between 1834 and 1837 they published Mappa Selenographica in four parts, which together made a whole-disk map 95 cm in diameter. Owing to the recent improvement in telescopes, their map surpassed all its predecessors. In 1838 they republished the map with a dissertation in their book Der Mond. It became the definitive work on the subject, but their convincing argument for the Moon being an airless and unchanging body prompted a hiatus in observing.

While reading Schroter’s 1791 Selenotopographische Fragmente as a boy, J. F.J. Schmidt in Germany decided to study the Moon. After acting as assistant at various German observatories, in 1858 he became director of the Athens Observatory and set out to make a full-disk map with a diameter of 180 cm which would show more craters, rilles and mountains than its predecessors. In fact, for many years he was the only observer engaged in systematic lunar work! He completed the observations in 1868, having produced in excess of 1,000 sketches, measured the positions of over 4,000 points, catalogued 278 rilles and used shadow details to measure the depths of craters and the heights of mountains. When issued in 25 sections in 1874, the map specified 33,000 craters and the heights of 3,000 mountains. In 1878 he reprinted it in his book Charte der Gebirge des Mondes. Having acquired Lohrmann’s files, Schmidt had his predecessor’s map engraved at the scale originally intended, and in 1878 published it as Mondkarte in 25 Sektionen; it would have been a fine map for its time, but was now obsolete.

In 1864 the British Association established a Lunar Committee, but this achieved little. In 1876 Edmund Neville Nevill (using the surname Neison) published a book, The Moon, which included a full-disk map in 22 sections with a diameter of 60 cm. In fact, it was a reworking of Beer and Madler’s chart supplemented with a detailed description of each named feature – some 500 in all – making it a monumental work. It stimulated sufficient interest to prompt the establishment of the Selenographical Society with William Radcliffe Birt, one of the most active of British amateur lunar observers of that time, as president, and Neison as secretary. But it was disbanded in 1883 after the death of Birt in 1881 and the departure of Neison in 1882 to become the first director of the Natal Observatory in Durban in South Africa. Nevertheless, when the British Astronomical Association was established in London in 1890 the former members of the Selenographical Society set up a Lunar Section as a means of coordinating their activities. T. G. Elger, the Director of the Lunar Section, published a book entitled The Moon in 1895 to assist new observers.

By 1890 photography had matured sufficiently to facilitate surveys of the Moon. Two photographic atlases were published in 1897. Atlas Photographique de la Lune covered the whole face in many small sections using plates taken by Maurice Loewy at the Paris Observatory and text provided by his assistant Pierre Puiseux. The Lick Observatory Atlas of the Moon by Edward Singleton Holden comprised 19 sheets of reproduced photographs. After establishing a temporary astronomical outstation in Jamaica, W. H. Pickering of the Harvard College Observatory set himself the task of photographing the lunar disk in several sections at five illumination phases. When The Moon – A summary of existing knowledge of our satellite, with a complete photographic atlas was published in 1903 it was the first true atlas, because the pictures were reproduced at the same scale. Although Jamaica had particularly clear skies, Pickering’s pictures of the Moon were still blurry and so there remained scope for visual studies, particularly in the limb regions – but as professional astronomers turned their attention to the stars and even more distant objects, they left the Moon, which they regarded as a source of ‘light pollution’, to their amateur brethren.

On the nights of 12 to 15 September 1919 Francis Pease photographed the Moon while testing the new 100-inch Hooker reflector at the Mount Wilson Observatory in California. Walter Goodacre was born in 1856, lived in London, and developed an interest in astronomy as a boy. In 1910, after making thousands of observations, he compiled a whole-disk map of the Moon almost 200 cm in diameter. In 1931, by which time he had replaced T. G. Elger as Director of the Lunar Section of the British Astronomical Association, he published The Moon. It featured a map which combined fine detail obtained by visual observing in good ‘seeing’, with positional accuracy derived from the Mount Wilson photographs, but few copies were issued.

The International Astronomical Union was established in 1919 to oversee general issues, and it took responsibility for regulating lunar nomenclature.

KENNEDY’S CHALLENGE

NASA was well placed to exploit the new administration’s willingness to expand the space program. Its long-term planning was impressive for its detail, in particular

This report was excerpted in the New York Times on 12 January 1961, and is sometimes wrongly dated as such.

because George Low’s committee had costed the accelerated plan – concluding that it would require $7 billion to land a man on the Moon by the end of the decade. In January 1961 Low briefed Keith Glennan on the forthcoming hearings for NASA’s budget, but Glennan expected to be replaced by the new administration and so was in a weak position.

On Johnson becoming Kennedy’s Vice President, Robert S. Kerr took over from him the chairmanship of the Senate Committee on Aeronautical and Space Sciences. After consulting Kerr, Johnson recommended James E. Webb to succeed Glennan as NASA administrator, and Webb took over on 14 February. Whereas Glennan was a scientific administrator with a conservative outlook, Webb was a political operator. He had served as Director of the Bureau of the Budget between 1946 and 1949 and Undersecretary of State from then until 1952 in the Truman administration. He had been a director of Kerr’s oil and uranium conglomerate, Kerr-McGee Oil Industries, and simultaneously a director of the McDonnell Aircraft Company.

Webb immediately set out to obtain the funding that was earlier denied for Apollo and the Saturn launch vehicle. When the Bureau of Budget refused, Webb wrote to Kennedy in early March that Eisenhower had “emasculated the 10-Year Plan before it was one year old’’, and if the funding were not made available it would “guarantee that the Russians will, for the next five to ten years, beat us to every exploratory space flight’’. To ram home the message in terms that Kennedy would appreciate, Webb said, “We have already felt the effects of the fact that they were the first to place a satellite into orbit, have intercepted the Moon, photographed the back side of the Moon, and have sent a large spacecraft to Venus. They can now orbit seven and a half ton vehicles about the Earth, compared to our two and a half tons, and they have successfully recovered animals from flights of as much as 24 hours. Their present position is one from which further substantial accomplishments can be expected, and our best information points to a steadily increasing pace of successful effort on a realistic timetable.’’

On 23 March Kennedy met with Lyndon Johnson, Jerome Wiesner, David Bell of the Bureau of Budget and Edward C. Welsh, a former aide to Johnson who was now serving as Executive Director of the National Aeronautics and Space Council, of which Johnson was chairman. Kennedy agreed to increase funding for the Saturn launch vehicle, but said he would need to deliberate further on the Apollo spacecraft – he would decide in the autumn, he said.

Just when NASA began to think that it might beat the Soviets to a manned space flight, on 12 April 1961 Yuri Alexseyevich Gagarin made a single orbit and landed safely. Webb told Congress, in budget hearings then underway, that NASA could certainly work faster if its funding was increased.

The next evening Kennedy met at the White House with Jerome Wiesner, David Bell, James Webb, Hugh Dryden, Theodore Sorensen, who was a friend and advisor, and Hugh Sidey, a journalist for Life magazine who was one of Kennedy’s friends, and put to them the question, “at what point we can overtake the Russians’’. NASA opened with a space station to be assembled in Earth orbit to serve as a jumping off point for a future mission to the Moon. But, it pointed out, if the Soviets were on the same plan they would likely remain in the lead for some considerable time. Kennedy

Гro.’ere Index

 

KENNEDY’S CHALLENGE
KENNEDY’S CHALLENGE
KENNEDY’S CHALLENGE
KENNEDY’S CHALLENGE

Deny

 

Reds Win Running Lead In Race To Control Space ^ 7:

 

T—

 

The Huntsville Times reports the first man in space.

 

image32

wanted to minimise this period, either by accelerating or by short circuiting the plan. Dryden said a ‘crash’ program might land a man on the Moon ahead of the Soviets, but it might cost as much as $40 billion. ‘‘The cost! That’s what gets me,’’ Kennedy mused. ‘‘When we know more, I can decide if it’s worth it or not. If somebody can just tell me how to catch up.’’ As the meeting broke up, Sorensen remained behind to discuss what had been said, and upon emerging told the others, ‘‘We’re going to the Moon!’’

On 19 April Kennedy summoned Johnson and told him he had decided to issue a momentous challenge. The next day, Kennedy sent a memo to Johnson seeking ‘‘an overall survey of where we stand in space’’. Specifically:

1. Do we have a chance of beating the Soviets by putting a laboratory in space, or by a trip around the moon, or by a rocket to land on the moon, or by a rocket to go to the moon and back with a man. Is there any other space program which promises dramatic results in which we could win?

2. How much additional would it cost?

3. Are we working 24 hours a day on existing programs. If not, why not? If not, will you make recommendations to me as to how work can be speeded up.

4. In building large boosters should we put [our] emphasis on nuclear, chemical or liquid fuel, or a combination of these three?

5. Are we making maximum effort? Are we achieving necessary results?

On 21 April Kennedy told reporters that his administration was considering the options and cost of space, and said, ‘‘If we can get to the Moon before the Russians, we should.’’

Johnson consulted NASA first, which said there was little chance of beating the Russians to a space station; it might be possible to beat them to lunar orbit; the best bet was a lunar landing. This matched Johnson’s thinking. NASA suggested 1967 as a target date because it was expected that the Soviets would attempt to make a lunar landing then in order to mark the 50th anniversary of the Bolshevik Revolution. As a result of the additional analysis by Low, the costing had been increased from the $7 billion estimate for a landing in 1969 to $22 billion; but a landing in 1967 would be $34 billion. Next, Johnson consulted the Pentagon, and the Air Force agreed that a manned lunar landing would be appropriate – even although the Air Force would not be allowed to perform it. Finally, Johnson consulted three businessmen whose judgement he trusted: Frank Stanton of the Columbia Broadcasting System; Donald Cook of the American Electric Power Service Corporation; and George Brown of Brown and Root, which was a construction company in Texas. The fact that none of them was involved in the aerospace industry that would be called upon to build the hardware for the program was a point in their favour, since it meant they were unbiased. At the National Aeronautics and Space Council on 24 April, Johnson, as Wiesner later described it, ‘‘went around the room saying, ‘We’ve got a terribly important decision to make. Shall we put a man on the Moon?’ And everybody said ‘yes’. And he said ‘Thank you’.’’

The scientific community was represented in the White House by Wiesner. The majority of space scientists were interested in particles and fields, and because this

Подпись: In accordance with our conversation I would like for you aa

Chairman of the Space Council to be in charge of making an overall survey of where we stand in apace.

1. Do we have a chance of beating the Soviets by putting a laboratory in space, or by a trip around the moon, pr by a rocket to land on the moon, or by a rocket to go to the moon and back with a man. Is there any other space program which promises dramatic results in which we could win?

2. How much additional would it cost?

3. Are we working 24 hours a day on existing programs. If not, why not? If not, will you make reconmenda – tions to me as to how work can be speeded up.

Д. In building large boosters should we put out

emphasis on nuclear, chemical or liquid fuel, or a confcination of these three?

5. Are we making maximum effort? Are we achieving necessary results?

Подпись: and other responsible officials to cooperate with you fully. I would appreciate a report on this at the earliest possible moment.

image33,image34

I have asked Jim Webb, Dr. Wiesner, Secretary McNamara

/в/ John F. Kennedy

The historic memo to Lyndon B. Johnson which led John F. Kennedy to challenge his nation to land a man on the Moon before the decade was out.

research did not require a human presence, money spent on sending men into space was by definition wasted. But Kennedy wanted “dramatic results” and the scientists were unable to offer this. To be fair, Kennedy invited Wiesner to suggest a terrestrial challenge that would serve the purpose, “… something with an overseas impact, like desalination or feeding the hungry”. However, Wiesner could see that the Moon was

shaping up to be the challenge, and advised the President “never to refer publicly to the Moon landing as a scientific enterprise”.

On 28 April Johnson submitted the National Aeronautics and Space Council’s recommendation:

Largely due to their concerted efforts and their earlier emphasis upon the development of large rocket engines, the Soviets are ahead of the United States in world prestige attained through impressive technological accomplishments in space. The US has greater resources than the USSR, etc. The country should be realistic and recognize that other nations, regardless of their appreciation of our idealistic values, will tend to align themselves with a country which they believe will be the world leader. The US can, if it will firm up its objectives and employ its resources, have a reasonable chance of attaining world leadership in space. If we don’t make a strong effort now, the time will soon be reached when the margin of control over space and other men’s minds through space accomplishment will have swung so far on the Russian side that we will not be able to catch up. Even in those areas in which the Soviets already have the capability to be first and are likely to improve upon such capability, the United States should make aggressive efforts, as the technological gains as well as the international rewards are essential steps in gaining leadership. Manned exploration of the Moon, for example, is not only an achievement with great national propaganda value, but is essential as an objective, whether or not we are first in its accomplishment – and we may be able to be first.

Kennedy was receptive to Johnson’s recommendation, but he postponed a formal decision until after the first manned Mercury mission, which came on Friday, 5 May 1961 when Al Shepard rode a Redstone missile on a suborbital arc.[19]

Over the weekend, Johnson met James Webb and Secretary of Defense Robert S. McNamara to draw up a formal recommendation to Kennedy’s memo of 20 April. Recommendations for our National Space Program: Changes, Policies and Goals, jointly authored by Webb and McNamara, said, “It is man, not merely machines, in space that captures the imagination of the world. All large-scale projects require the mobilization of resources on a national scale. They require the development and successful application of the most advanced technologies. Dramatic achievements in space, therefore, symbolize the technological power and organizing capacity of a nation. It is for reasons such as these that major achievements in space contribute to

KENNEDY’S CHALLENGE

image35

On 5 May 1961 a Redstone rocket lifts off with a Mercury capsule containing Alan B. Shepard for a suborbital mission.

 

image36

After his successful Mercury flight, Alan B. Shepard shakes hands with John F. Kennedy at the White House.

 

Подпись: 80 The Apollo commitment

In a speech to Congress on 25 May 1961 John F. Kennedy challenged his nation to land a man on the Moon before the decade was out.

image37national prestige.” They wrote, “even though the scientific, commercial or military value of [such an] undertaking may by ordinary standards be marginal or economically unjustified”, it nevertheless generated “national prestige”, which had value in its own right. Furthermore, “The non-military, non-commercial, non­scientific but ‘civilian’ projects such as lunar and planetary exploration are, in this sense, part of the battle along the fluid front of the Cold War.’’ This echoed Kennedy’s criticism of Eisenhower: whereas Eisenhower had been conscious of the cost and dismissive of national prestige, to Kennedy national prestige was the issue and the cost was secondary.

On 25 May Kennedy gave a speech to a joint session of Congress on the theme of Urgent National Needs. In view of recent space achievements by the Soviets, he proclaimed, ‘‘Now it is time to take longer strides, time for a great new American enterprise, time for this nation to take a clearly leading role in space achievement, which in many ways may hold the key to our future on Earth.’’ Having outlined the political background, he laid down the gauntlet. ‘‘I believe that this nation should commit itself to achieving the goal, before this decade is out, of landing a man on the Moon, and returning him, safely, to the Earth.’’ He had opted for a lunar landing precisely because it posed a great technical challenge. By literally ‘shooting for the Moon’, he was betting that America would not only catch up with the Soviet Union in space, but forge ahead. Having concluded that space was the arena of superpower politics, he was challenging his rival, Nikita Khrushchev, for world leadership. He had imposed the deadline to ensure that reaching the Moon was perceived as a race. He was also well aware of the magnitude of the task. ‘‘No single space project in this

period will be more impressive to mankind, or more important for the long-range exploration of space; and none will be so difficult or expensive to accomplish.” The sending of a man to the Moon was to be the modern form of the ancient practice of ‘single combat’, whereby opposing armies lined up and each dispatched a single warrior to decide the issue. To indicate that it was a matter of national honour, he added, ‘‘In a very real sense, it will not be one man going to the Moon; if we make this judgment affirmatively it will be an entire nation, for all of us must work to put him there.’’ And in order to emphasise what was at stake, he warned, ‘‘If we are to go only halfway, or reduce our sights in the face of difficulty, in my judgment it would be better not to go at all.’’

For Kennedy the Moon was a symbol and, in terms of what he wished to achieve it was an excellent symbol. He had the impression that the applause in Congress was ‘‘something less than enthusiastic”, as he told Sorensen immediately after giving the speech. But Johnson had read the mood well: there was only minor opposition in the House of Representatives, and the debate in the Senate lasted less than an hour – only five of the 96 senators spoke, and the floor was dominated by Robert Kerr, who was Johnson’s man. NASA’s budget was doubled without a formal vote being taken.

Developing Lunar Orbiter

CHANGING HORSES

In January 1961 the Hughes Aircraft Company was selected to build the Surveyor spacecraft. With a planned mass of about 1,125 kg at translunar injection, it would require the Atlas-Centaur. The plan was for the orbital version to provide wide-area mapping and reconnaissance of potential landing sites for the surface Surveyors and, later, for Apollo. The mass at touchdown was expected to be about 340 kg, of which 114 kg would be scientific instruments which would not only transmit pictures but also provide data on the physical, chemical, mineralogical and biological properties of the surface material. The initial schedule called for the first flight in 1964. It was also envisaged that as the project matured, an orbital variant would be equipped to serve as a communications relay for landers investigating sites on the far-side of the Moon.

On 23 March the Lunar Science Subcommittee at the Office of Space Sciences recommended that the orbiter have a TV system which was capable of providing: (1) full coverage of the limb areas (highly foreshortened to terrestrial observers) and of the far-side at a resolution of 1 km, (2) wide-area reconnaissance at a resolution of 100 metres, and (3) stereoscopic pairs of selected areas with sufficient resolution to discern objects 10 metres in size. On 5 December Charles Sonett, Chief of Lunar and Planetary Sciences at the Office of Space Sciences, asked William Cunningham to determine the status of the orbiter. On 12 January 1962 Cunningham reported that JPL would not be able simply to adapt the vidicon system developed for Ranger; a new TV system would be required, the development of which had not yet begun.

NASA called for JPL to define the design requirements for the Surveyor orbiter, maximising its commonality with the lander, by 1 September 1962, but owing to the problems the laboratory was facing with Ranger the orbiter received little attention.

Meanwhile, on 15 June 1962 the Office of Manned Space Flight compiled a list of the data that it required the Office of Space Sciences to supply on the environment around the Moon and on its surface – i. e. Brainerd Holmes’s Requirements for Data in Support of Apollo. In view of the urgency to feed such information into the design

of the Apollo vehicles, the Office of Space Sciences asked JPL whether Ranger could serve as the basis for an orbiter. JPL in turn asked the Hughes Aircraft Company to consider the possibility of a 360-kg orbiter which could be launched by the Atlas – Agena. Hughes replied that in order to meet this mass limit, the scientific payload could not exceed 27 kg, which was unrealistic in view of the activity to be pursued. JPL calculated that if the solid-rocket motor that Surveyor was to use in the initial phase of its descent to the Moon were to be used to augment the Agena in the translunar injection, it would be possible to increase the scientific payload to 57 kg, but this was still too little. Even although the development of the Centaur stage was running behind schedule, the Office of Space Sciences decided to proceed with the Centaur-based orbiter.[26] To meet the Apollo requirements, the orbiter would require to provide photography of potential landing sites capable of revealing protuberances and pits as small as 1 metre in size and slopes as shallow as 7 degrees. But even the stereoscopic views from the Surveyor orbiter’s TV system would have a resolution no better than 10 metres. A photographic system employing film would be needed to meet the requirements of the Office of Manned Space Flight.

On 21 September 1962 Oran Nicks, Director of Lunar and Planetary Programs in the Office of Space Sciences, asked Lee R. Scherer to form a committee to evaluate proposals which had been submitted by the Space Technology Laboratories and the Radio Corporation of America for a ‘lightweight’ lunar orbiter compatible with the Atlas-Agena.

On 23 October Joseph Shea, the Deputy Director of the Office of Manned Space Flight, specified the relative priorities of the data that Apollo would require from the Office of Space Sciences. There was a greater need for the information which a soft – lander would provide, since this would feed into the design of the Apollo vehicles, whereas the information from an orbiter would not be required until later, in mission planning. Shea stressed that if funding was tight in Fiscal Year 1963, then the Office of Space Sciences should favour the lander over the orbiter.

Scherer reported to Nicks on the issue of an Agena-based orbiter on 25 October. The proposal by the Radio Corporation of America was for a Ranger bus to make a lunar flyby, dropping off a 200-kg package which would insert itself into orbit. The orbiter would be 3-axis stabilised and use a vidicon system (no doubt a development of the camera the company had provided for Ranger Block II) to provide pictures at a resolution of 130 metres in the wide-angle coverage and 30 metres in the narrow – angle coverage. The Space Technology Laboratories had envisaged an orbiter with a mass of 320 kg. It would have a monopropellant engine which was capable of firing several times. In addition to a midcourse manoeuvre and orbit insertion, this engine would permit changes to the orbit. One mission profile would be to enter a circular polar orbit at an altitude of 1,600 km and map the entire Moon, resolving objects as small as 18 metres in size. Alternatively, it could be placed into equatorial orbit at an altitude of 40 km to photograph that zone with a resolution of 0.5 metre. It would be spin-stabilised, and use a ‘spin scan’ camera of a design similar to that proposed by the RAND Corporation in 1958. It would use film to obtain a higher resolution than was obtainable using a vidicon. Scherer reported that only the proposal by the Space Technology Laboratories offered the prospect of meeting the requirements set by the Office of Manned Space Flight for imaging resolution, and he recommended that the company further refine the concept so as to enable the Office of Space Sciences ‘‘to establish the confidence needed [to consider] a flight program of this type, should it be deemed preferable to a Centaur-based orbiter’’. In fact, once the viability of an Agena-based reconnaissance orbiter had been established, this in itself undermined the case for pursuing the Surveyor orbiter.

On 26 October, Clifford Cummings, unaware of Scherer’s study, wrote to advise Oran Nicks that JPL was about to conduct a study to refine the configuration of the Surveyor orbiter in order to specify how it would perform its mission. In his reply on 8 November, Nicks pointed out that the Office of Space Sciences was looking into the possibility of an Agena-based orbiter.

On 2 January 1963 Nicks asked Floyd L. Thompson, the Director of the Langley Research Center, to consider the possibility of his staff taking on the development of a lightweight orbiter. Thompson set up an internal feasibility study. After the Space Technology Laboratories had refined its concept, a review was held at Langley on 25 February involving representatives of the company, the Office of Space Sciences, the Office of Manned Space Flight, Langley and Bellcomm. Lee Scherer and Gene Shoemaker reported on a study they had undertaken for Nicks to determine how a lightweight orbiter might satisfy the photographic requirements of Apollo. Dennis Jones of Bellcomm reported an assessment made for Shea on the degree to which an orbiter might support the manned and unmanned exploration of the Moon. A second meeting on 5 March agreed that not only was a lightweight orbiter viable, it would also significantly support Apollo. Langley then sent a delegation headed by Clinton E. Brown to brief Robert Seamans and present the case for Langley taking on such a project; Seamans authorised planning to proceed.

In order to assist Langley draw up the request for proposals, in April 1963 the Office of Manned Space Flight refined its requirements. The critical needs were: (1) data on the radiation flux in lunar space over a typical 2-week period; (2) a summary and analysis of all efforts for short-term prediction of severe solar proton events; (3) measurements of particles capable of penetrating 0.01 cm and 0.1 cm of aluminium in an average peak 2-week period of micrometeoroid activity; and (4) photographic data capable of showing protuberances 3.5 metres tall and slopes of 15 degrees in an area of the lunar surface with a radius of 60 metres (to be provided by the autumn of 1965) and then equivalent data showing 50-cm protuberances and 8-degree slopes in an area with a radius of 1,600 metres. Other needs were: (1) measurements of the distribution of slopes greater than 15 degrees in areas of 3.5 metres radius; and (2) the greatest possible coverage of the zone within 5 degrees of the lunar equator with a resolution of 25 metres or better.

On 25 April Edgar Cortright put it to Homer Newell that since one successful orbiter could be worth “dozens of successful Ranger TV impactors”, the three new Rangers which had recently been funded in order to obtain high-resolution pictures and gamma-ray and radar reflectance data on the Moon should be cancelled. Newell accepted this reasoning and passed the recommendation to Robert Seamans, who concurred on 12 July. Later in the year, the second batch of rough landing Rangers was also cancelled.