Category Escaping the Bonds of Earth

OUTSIDER

When Bill Anders joined NASA in October 1963, he stood out: among the Fourteen, six astronaut candidates lacked test-piloting credentials, and he was one of these apparent ‘outsiders’. As a child, he had never had aspirations of becoming a test pilot, but rather wanted to follow in the footsteps of his father, Arthur Anders, and become a career naval officer. William Alison Anders was born in Hong Kong on 17 October 1933 and, in his youth, was an active Boy Scout, receiving its second-highest rank. He earned a bachelor’s degree from the Naval Academy in 1955, but opted for a commission in the Air Force, serving as a fighter pilot in an all-weather interceptor squadron in Iceland.

Anders’ decision to move away from naval service was made in part by the extraordinary number of fatal accidents he saw as a midshipman: one aircraft, for example, landed on the carrier, missed the arresting net, hit a line of parked jets, careered off the deck and plunged into the sea. Anders accepted the risks of his new­found love of aviation, but preferred to face such risks in air-to-air combat, rather than whilst attempting to land. After Iceland, Anders’ next step, with 1,500 hours in his flight logbook, was to apply for test pilot school. He was rejected, on the basis that the school was ‘‘pushing academics’’ and desired him to earn an advanced degree.

With test pilot school still at the back of his mind, Anders enrolled in the Air Force Institute of Technology at Wright-Patterson Air Force Base in Dayton, Ohio, graduating with a master’s degree in nuclear engineering in 1962. He had wanted to study astronautical engineering, but places were full, although Anders simulta­neously took night school classes in aeronautics at Ohio State University. Upon receipt of his master’s qualification, Anders again tried for test pilot school, only to learn that it was not recruiting students, and he moved instead to become a T-33 instructor pilot at Kirtland Air Force Base in Albuquerque, New Mexico.

The following year, 1963, proved life-changing for Anders. He again applied for test pilot school, only to learn that now, under the new commandant, Chuck Yeager, it was looking for candidates with more flying experience. Unperturbed, he submitted his application and in June, whilst waiting to hear of the outcome, he learned that NASA were recruiting for its new class of astronauts. Test piloting qualifications, it turned out, were no longer mandatory and Anders had everything that the space agency needed: he was younger than 35, had an advanced degree and his logbook had now expanded to more than the 2,000 required hours.

When he was called to interview, Anders stressed his nuclear engineering work, aware that flights to the Moon would surely require an understanding of radiation hazards in cislunar space. That October, he was picked as one of the Fourteen… and, ironically, received a letter rejecting him from test pilot school! In his early years as an astronaut, Anders supervised Apollo’s environmental controls and his performance as the capcom during the Gemini VIII emergency in March 1966 quite possibly contributed to his assignment as Neil Armstrong’s pilot on the Gemini XI backup crew. By the end of that year, Anders had drawn his first actual flight assignment: to Frank Borman’s E mission, during which he hoped to get the opportunity to test-fly the lunar module in high Earth orbit. All that changed in August 1968 when Anders became a part of history: one of the first human explorers to visit the Moon.

FIRST AMONG EQUALS

In a strange kind of way, Neil Armstrong’s work during his first couple of years as an astronaut had helped pave the way for the selection of Dave Scott, his 33-year-old colleague aboard Gemini VIII. Deke Slayton had given Armstrong responsibility for mission operations and training and entrusted him to devise a system which would identify how many astronauts would be needed at any given time. It was this schematic, wrote James Hansen in his biography of Armstrong, that ‘‘allowed Slayton to determine when additional astronauts needed to be brought in. . . culminating in Houston’s announcement in June 1963 that NASA was looking for a new class of astronauts’’.

Slayton’s decision, he recounted, ‘‘was based on planning documents that were starting to arrive from [NASA] Headquarters’’. The 1962 astronaut intake, together with the remaining members of the Mercury Seven who were still on flight status, barely provided enough seats for Project Gemini. Early predictions for Apollo envisaged at least 12 missions in Earth and lunar orbit before a landing on the Moon could be attempted. That required Slayton to accept ‘‘a minimum of ten new astronauts in 1963 and… as many as twenty”. Fourteen newbies duly arrived in Houston in October of that year.

Although the selection criteria changed slightly – a more rigid upper age limit of 34 and the elimination of the need for the new astronauts to be test pilots being the main differences – David Randolph Scott still fitted perfectly into the classic mould of a spacefarer. Tall, athletic and with a middle name honouring the Air Force base on which he was born, it has often been said that it was no accident that Scott was the first of the 1963 group of astronauts to fly into space… and eventually would become the first of his class to command a mission. “In some circles,” noted Andy Chaikin in his book about Project Apollo, “there was a joke that if NASA ever came out with an astronaut recruiting poster, Dave Scott should be on it.” Further, Chaikin added, even astronauts who did not get on with Scott placed him at the top.

Despite the near-disaster which befell it, Gemini VIII marked the only mission in which the entire crew would one day set foot on the Moon; Armstrong as the first man to tread its dusty surface at the Sea of Tranquility and Scott in command of Apollo 15 in July 1971.

Scott was born on Randolph Air Force Base in San Antonio, Texas, on 6 June 1932, the son of an Air Force officer and progeny of a strict, frugal military family who instilled in him the virtues of personal discipline and devotion to setting and achieving ambitious life goals. In his autobiography, jointly co-authored with cosmonaut Alexei Leonov, Scott recalled watching Jenny biplanes soaring over Randolph as a three-year-old boy and was fascinated that aboard one of them was his father, Army Air Corps flier Tom Scott. From that tender age, the young Scott set his sights on someday becoming a pilot.

With a military father, the family moved many times during his childhood, from Texas to Indiana, abroad to the Philippines and, in 1939, back to the United States. Scott’s father was posted overseas after Pearl Harbour to support the war effort and the young boy developed an interest in model aircraft. Then, when his father returned at the end of the conflict, Scott received his first flying lesson.

Despite a desire to study at West Point, Scott won a swimming scholarship to read mechanical engineering at the University of Michigan in 1949. After a year in Michigan, he was summoned for a physical at West Point, passed and headed to upstate New York to begin preparations for a military career. In his autobiography, Scott would credit his four years at West Point – plus his own upbringing – as “the most valuable and formative… of my life’’.

Ultimately, in 1954, he graduated fifth in his class and opted to join the Air Force, receiving initial flight training at Marana Air Force Base in Tucson, Arizona. He subsequently moved to Texas to begin working on high-performance jets, followed by gunnery preparation and assignment to a fighter squadron in Utrecht, Holland. Whilst in Europe, Scott flew F-86 Sabre and F-100 Super Sabre jets under a variety of weather conditions and, in October 1956, when Soviet tanks rolled into Budapest, his squadron was placed on high alert for the first time.

Three years later, as the Mercury Seven were introduced to the world’s media, Scott watched from afar with scepticism; why, he wondered, were they abandoning

such promising military careers? His own focus was upon gaining an advanced degree in aeronautics from Massachusetts Institute of Technology (MIT) and achieving admission to test pilot school. His work at MIT, he later wrote, “was like trying to drink water from a high-pressure fire hydrant… Compared to the hard grind of MIT, the five or six years I had spent flying fighter jets felt like playing.” He and his wife, Lurton, also had a newborn daughter, Tracy.

As part of his master’s degree, Scott was introduced to the new field of ‘astronautics’ – “my first exposure to space’’ – and his dissertation focused upon the mathematical application of guidance techniques and celestial navigation. This undoubtedly proved of benefit during Projects Gemini and Apollo, both of which depended heavily upon rendezvous and docking. Scott passed his final exams shortly after John Glenn’s Friendship 7 mission, hoping for reassignment to test pilot school… only to be detailed instead as a professor of aeronautics and astronautics at the Air Force Academy. Fortunately, a conversation with a sympathetic superior officer led to a change of orders and Scott reported to the Experimental Test Pilot School at Edwards Air Force Base in California.

Graduation was followed, in mid-1963, by the lengthy application process to join NASA’s astronaut corps. In his autobiography, Scott recalled undergoing cardiograms, running on treadmills and enduring hypoxia evaluations, in which he was starved of oxygen to assess his physical response. The psychologists were especially difficult to please. When asked about his MIT days and how he had liked Boston, Scott replied that he found New Englanders cold, aloof and ‘‘a little hard to get to know’’ … only to discover that the stone-faced psychologist was a born-and – raised Bostoner!

It obviously had little adverse impact on his application, however, and in October 1963 Scott and 13 other candidates – Edwin ‘Buzz’ Aldrin, Bill Anders, Charlie Bassett, Al Bean, Gene Cernan, Roger Chaffee, Mike Collins, Walt Cunningham, Donn Eisele, Ted Freeman, Dick Gordon, Rusty Schweickart and Clifton ‘C. C.’ Williams – were notified of their assignment to NASA. Of these, eight were test pilots, whom Slayton intended to use ‘‘for the more immediately difficult work’’ as solo command module pilots for the Apollo missions, while the others represented a mixture of operational military fliers, engineers or researchers. The latter, added Slayton, ‘‘would get their chance, too, but on the development end of things’’. The excitement of the accelerating lunar effort, Scott wrote, was palpable, even in the wake of President John Kennedy’s assassination a few weeks later.

The new astronauts worked together surprisingly well, with the Air Force pilots kidding their Navy counterparts that they couldn’t bring their jets down without a thump. In return, the naval aviators retorted that the Air Force fliers needed far too much runway on which to land. Despite the steadily burgeoning number of astronauts, they were still offered deals on cars, small Life magazine contracts and ‘‘the banks all wanted to have our accounts’’. Unlike the Mercury Seven, however, their privacy was better respected and their wives and children were pestered less by journalists from other publications.

After initial training, Scott was assigned guidance and navigation as his area of responsibility and in June 1965 served as one of the backup capcoms for Gemini IV and America’s first spacewalk. Less than three months later, at the end of August, Deke Slayton caught up with him for “a word’’. Now that Gemini V was over, Slayton said, Neil Armstrong was free from his backup duties and would be teamed with Scott for Gemini VIII, scheduled for March 1966. The mission, Slayton added, would involve a lengthy, two-hour EVA for Scott. The boy from Randolph Air Force Base could not have been more delighted.

DR. RENDEZVOUS

Not long after his return from Gemini IX-A, astronaut Gene Cernan was summoned to Deke Slayton’s office and posed an unusual question.

‘‘Geno, how soon can you be ready to fly again?’’

‘‘Just say the word, Deke. When?’’

“Right now. Would you be willing to jump from backup to prime? Fly [Gemini] XII with Lovell?”

The year 1966 had certainly been a dramatic one for Cernan. When it began, he and Tom Stafford confidently looked forward to flying Gemini XII – the last manned mission in the series – themselves. Then, with awful suddenness, the deaths of Elliot See and Charlie Bassett in February pushed them from backup to prime crew on Gemini IX. Following his return from his first spaceflight, Cernan had been given a ‘dead-end’ slot, with Gordo Cooper, to back up Jim Lovell and Buzz Aldrin on Gemini XII. Now, with barely two-thirds of the year gone, Slayton was offering to break his own crew-rotation system, bumping Aldrin from the mission. Cernan’s first question to Slayton was a simple one. Why?

The reason was the Astronaut Manoeuvring Unit (AMU) – the Air Force-built rocket armchair which Cernan was originally detailed to test in June 1966 – whose military sponsors were pushing strongly to fly again on Gemini’s last mission. Without giving much away, Slayton told Cernan simply that he was the best man to fly the AMU, which was probably true, but a number of contributory factors centred on Buzz Aldrin himself: a man of mathematical and engineering genius, the first astronaut to possess a doctorate, an unquestioned expert in the field of space rendezvous. . . and a constant worry to Slayton. Aldrin had already raised eyebrows during Gemini IX-A, specifically those of Bob Gilruth and Chris Kraft, when he advocated having Cernan cut the lanyards of The Blob. Although not an outrageous suggestion, Slayton acquiesced, Aldrin’s advice was a little too adventurous in light of NASA’s limited EVA expertise.

Gilruth, Kraft and Slayton were not the only ones with worries about Aldrin. In his autobiography, Cernan hinted strongly that Aldrin’s intelligence was tempered by a seeming inability to stick to one topic: he had a tendency to fly off at tangents and drastically re-engineer everything, at a time when NASA had little time to do so. Astronauts and their wives would roll their eyes when Aldrin collared them, even over coffee, and engaged them in hours-long discussions of the intricacies of celestial navigation and mechanics. Coupled with reports of his performance in the Gemini simulators, it was Slayton’s judgement that the AMU test flight should be entrusted to Cernan, rather than Aldrin.

In his own defence, Aldrin would blame the decision on problems experienced by both Cernan and Dick Gordon on their EVAs: exhaustion, fogged-up visors and a difficulty in performing even simple tasks. ‘‘An urgent meeting of senior officials concerned with the Gemini XII EVA,’’ Aldrin wrote, ‘‘was held at the end of September and… they decided arbitrarily that I stood a poor chance of putting the innovative AMU backpack to good use. They felt the risks outweighed the benefits.’’

Despite the risks to his colleague’s career, Cernan accepted Slayton’s invitation on the spot – ‘‘when Deke asked if you would take a mission, there was only one answer’’ – and would have flown Gemini XII had not the decision been made that an AMU test was too risky. Gemini XII’s EVAs would focus instead on less dramatic evaluations of a spacewalker’s performance outside the pressurised confines of his spacecraft. Edwin Eugene Aldrin Jr, nicknamed ‘Dr Rendezvous’ behind his back, retained his place on the mission. Born in Glen Ridge, New Jersey on 20 January 1930, the son of an Army Air Corps pilot and a mother whose maiden name, propitiously, happened to be ‘Moon’, Aldrin’s development, even into adulthood, was very much guided by his father.

Naturally, in light of his father’s career, the man who would someday fly Gemini XII and become the second person to walk on the Moon was brought up with aviation in his blood. He first flew aboard an aircraft with his father in 1932, when he was barely two years old. (As a child, he earned the nickname ‘Buzz’ from his young sister, who, unable to pronounce ‘brother’, called him her ‘buzzer’.) Graduation was followed by enrolment in a military ‘poop school’ – aimed at preparing him for the Naval Academy at Annapolis – although Aldrin sought the Military Academy at West Point. Despite his father’s outspoken preference for the Navy, which he considered ‘‘took care of its people better’’, his son persisted and eventually won his reluctant approval.

When Aldrin graduated third in his class from West Point in 1951, his father’s immediate reaction was a question: who had finished first and second? He was not accepted for a coveted Rhodes postgraduate scholarship and instead entered the Air Force, earning his pilot’s wings later that same year after initial training in Bryan, Texas. During the conflict in Korea, Aldrin was attached to the 51st Fighter Wing, flying F-86 Sabres, and by the time hostilities ended in the summer of 1953 he had no fewer than 66 combat missions in his military logbook. Just a month before the end of the war, one of Aldrin’s gun-camera photographs – a Russian pilot ejecting from his stricken MiG – ended up in Life magazine.

In total, Aldrin returned from Korea having shot down three MiGs. Back in the United States, he became a gunnery instructor at Nellis Air Force Base in Nevada and in 1955 was accepted into Squadron Officer School in Montgomery, Alabama. At around the same time, he met and married Joan Archer and shortly thereafter became the father of a son, James. Professionally, his military career prospered: he was assigned as an aide to General Don Zimmerman, the dean of the new Air Force Academy, then moved to Germany in 1956, flying the F-100 Super Sabre as part of the 36th Fighter-Day Wing, stationed in Bitburg. During this time, he became a father twice more: to Janice and Andrew.

Before pursuing his next ambition of test pilot school, Aldrin, like his Bitburg flying comrade Ed White, sought to gain further education and was accepted into the Massachusetts Institute of Technology on military detachment for a doctorate of science degree in astronautics. His 259-page ScD thesis, completed in 1963, just months before his selection as an astronaut, was entitled ‘Line of Sight Guidance Techniques for Manned Orbital Rendezvous’. He chose the topic, he later wrote, because he felt it would have practical applications for the Air Force and aeronautics, although it also drew the attention of NASA, which was by now looking at lunar-orbital rendezvous for its Apollo effort.

Aldrin dedicated his thesis to future efforts in human exploration, wistfully remarking ‘‘if only I could join them in their exciting endeavours’’. By this time, of course, his application for the 1963 astronaut class was already being processed; he had tried to gain admission the previous year, seeking a waiver for his lack of test – piloting experience, but this time achieved success. A concern about his liver function, thanks to a bout of infectious hepatitis, did not prevent Aldrin from becoming one of the 14 astronauts named to the world that October.

Assigned to work on mission planning, his early days saw him focusing his attention on rendezvous and re-entry techniques… and, gradually, as each Gemini crew was named, he became increasingly frustrated that he was receiving no flight assignment. At one stage, he even approached Deke Slayton to stress his confidence in his own abilities – that his qualifications and understanding of orbital rendezvous far exceeded those of anyone else in the office – and was politely told that his comments would be noted.

Shortly thereafter, in early 1966, Aldrin and Jim Lovell were assigned as the backup team for Gemini X. His heart sank. Taking into account Slayton’s three – flight rotation system for backup-to-prime crews, there would be no Gemini XIII to which Aldrin and Lovell could aspire. It was, in effect, a ‘dead-end’ assignment. “Apparently, petitioning Deke – an arrogant gesture by ‘Dr Rendezvous’ – had not been well-received by the stick-and-rudder guys in the Astronaut Office,’’ Aldrin wrote. ‘‘By being direct and honest rather than political, I’d shafted myself.’’ All that changed on the last day of February, when See and Bassett were killed and their Gemini IX backups were pushed into prime position. In mid-March, Lovell and Aldrin were named as the new Gemini IX backups, with a formally-unannounced (but anticipated) future assignment as the prime Gemini XII crew.

For Aldrin, whose Nassau Bay backyard bordered that of the Bassetts, it was a devastating way to receive his long-desired flight assignment. Three weeks after the accident, he and Joan visited Jeannie Bassett to tell her the news. ‘‘I felt terrible,’’ he wrote, ‘‘as if I had somehow robbed Charlie Bassett of an honour he deserved.’’ Jeannie responded with quiet dignity and characteristic grace: her husband, she explained, felt that Aldrin ‘‘should have been on that flight all along. . . I know he’d be pleased’’.

C-PRIME

That August, around the time that Anders’ E mission was beginning its metamorphosis into C-prime, Apollo command module 103 arrived at Cape Kennedy for testing. Its mission, unofficially, ranged from circumlunar to fully orbital, with around ten circuits of the Moon planned. During the translunar coast, to qualify the ‘make-or-break’ SPS engine, it would be test-fired for a few seconds. If the engine refused to work, the astronauts could still be brought home safely, thanks to a safety feature built into Apollo 8’s trajectory design. Known as the ‘free return’, it would allow the crew to essentially loop around the Moon and use its gravitational influence to ‘slingshot’ them back to Earth without using the SPS. In fact, if Borman, Lovell and Anders did find themselves with a useless engine, they would only need to perform a couple of mid-course correction burns, using the service module’s thruster quads, to keep them on track for home.

Aside from the chance of an SPS failure, a host of other concerns worried Borman. One of them surrounded Apollo 8’s splashdown in the Pacific at the end of the six-day mission. To achieve a splashdown in daylight hours would require a trajectory design which included at least 12 lunar orbits. Borman, though, could not care less whether he landed in daylight or darkness. ‘‘Frank didn’t want to spend any more time in lunar orbit than was absolutely necessary,’’ wrote Deke Slayton, ‘‘and pushed for – and got – approval of a splashdown in the early morning, before dawn.’’ Apollo 8 would stick at ten orbits. To understand Borman’s reluctance to do more than was necessary is to understand part of his character and military bearing: he was wholly committed to The Mission, whatever it might be. On Gemini VII, he strenuously rejected any addition to the flight plan which might complicate his and Lovell’s chances of fulfilling their primary objective: to spend 14 days in space. Now, on Apollo 8, Borman’s mission was to reach the Moon and bring his crew home safely. Nothing else mattered.

All non-essential, ‘irrelevant’ requests irritated him. “Some idiot had the idea that on the way to the Moon, we’d do an EVA,’’ he recounted years later in a NASA oral history. ‘‘What do you want to do? What’s the main objective? The main objective was to go to the Moon, do enough orbits so that they could do the tracking, be the pathfinders for Apollo 11 and get your ass home. Why complicate it?’’

The four months leading up to the mission were conducted at a break-neck pace. The lunar launch window opened on 21 December, at which time Mare Tranquilitatis (the Sea of Tranquility, a low, relatively flat plain tipped as a possible first landing site) would be experiencing lunar sunrise and its landscape would be thrown into stark relief, allowing Borman, Lovell and Anders to photograph and analyse it. In the final six weeks before launch, the Apollo 8 crew regularly put in ten-hour workdays, with weekends existing only to wade through piles of mail. At the end of November, outgoing President Lyndon Johnson threw them a ‘bon voyage’ party in Washington. Then, on the evening of 20 December, the legendary Charles Lindbergh, first to fly solo across the Atlantic, visited their quarters at Cape Kennedy. During their meal, the topic of conversation turned to the Saturn V rocket, which would burn over 18,000 kg of fuel in its first second of firing. Lindbergh was astounded. ‘‘In the first second of your flight tomorrow,’’ he told them, ‘‘you’ll burn ten times more fuel than I did all the way to Paris!’’

Shortly after 2:30 am on launch morning, Deke Slayton woke them in Cape Kennedy’s crew quarters and joined them for the ritual breakfast of steak and eggs. Also in attendance were chief astronaut Al Shepard and Apollo 8 backups Neil Armstrong and Buzz Aldrin. (The third backup crew member, Fred Haise, was busily setting switch positions inside the command module at Pad 39A.) Shortly thereafter, clad in their snow-white space suits and bubble helmets, they arrived at the brilliantly-floodlit pad, where their Saturn V awaited. First Borman, then Anders and finally Lovell took their seats in the command module, joining Haise, who had by now finished his job of checking switches. After offering them his hand in solidarity and farewell, Haise crawled out of the cabin and the heavy unified hatch slammed shut at 5:34 am. Years later, Bill Anders would tell Andrew Chaikin that, at one point, he glanced over at a window in the boost-protective cover and saw a hornet fluttering around outside. ‘‘She’s building a nest,’’ he thought, ‘‘and did she pick the wrong place to build it!’’

As their 7:51 am launch time drew closer, a sense of unreal calm pervaded Apollo 8’s cabin. With five minutes to go, the white room and its access arm rotated away from the spacecraft and, shortly thereafter, the launch pad’s automatic sequencer took charge of the countdown, monitoring the final topping-off of propellants needed by the Saturn to reach space. Sixty seconds before launch, the giant rocket was declared fully pressurised and it transferred its systems to internal control. As the countdown ticked into the final dozen seconds, Borman, Lovell and Anders, despite being cocooned inside their space suits, could faintly hear the sound of fuel pouring into the combustion chambers of the five F-l engines, a hundred and ten metres below. As the clock inside the command module read ‘T-3 seconds’, that faint sound was replaced by a distant, thunder-like rumbling and, at some point in the calamitous commotion that followed, the first Saturn V ever to be trusted with human passengers took flight.

‘‘Liftoff,’’ radioed Borman, gazing at the clock on his instrument panel. ‘‘The clock is running.’’ After the mission, all three men would have their own recollections of what it was like to launch atop the biggest and most powerful rocket ever built, but Chaikin summed it up best when he quoted Bill Anders: they felt as if they were little more than helpless prey in the mouth of a giant, angry dog.

Forty seconds into the climb, the rocket broke through the sound barrier and G loads on the three astronauts climbed steadily – three, now four, and still climbing – but when they hit 4.5 the uncomfortable feeling of intense acceleration ended as the Saturn’s S-IC first stage burned out and separated. ‘‘The staging,’’ Borman recounted, ‘‘from the first to the second stage, as we went from S-IC cutoff to S-II ignition, was a violent manoeuvre: we were thrown forward against our straps and smashed back into the seat.’’ So violent, in fact, was the motion that Anders felt he was being hurled headlong into the instrument panel. Seconds later, the now – unneeded escape tower and the command module’s boost-protective cover were jettisoned, flooding the cabin with daylight as windows were uncovered. For Anders, his first glimpse of Earth from space – mesmerising clouds, vivid blue ocean and a steadily darkening sky – were electrifying.

A little under nine minutes after launch, the S-II finally expired and the S-IVB picked up the remainder of the thrust needed to achieve orbit. ‘‘The smoothest ride in the world’’ was how Borman would later describe riding the Saturn’s restartable third stage, before it, too, shut down, at 8:02 am. Barely ll minutes had passed since leaving Cape Kennedy and the Apollo 8 astronauts were in orbit. In less than three hours’ time, assuming that their spacecraft checked out satisfactorily, they would relight the S-IVB for six minutes to begin the translunar injection, or TLI, burn and set themselves on course for the Moon. However, if Apollo 8 did not pass its tests with flying colours and the lunar shot was called off, they would be consigned to what had been uninspiringly termed ‘the alternate mission’: an Apollo 7-type jaunt for ten long days in Earth orbit, with little to do. Borman could think of nothing worse.

Indeed, at one stage, Lovell, working under one of the couches to adjust a valve, accidentally inflated his space suit’s life vest and his commander gave him a dirty look. In true Frank Borman fashion, nothing would be permitted to interfere with or distract their attention from The Mission. At length, it was Capcom Mike Collins, who had been recovered from his neck surgery since early November and had even fruitlessly approached Deke Slayton with a view to staying on the crew, who gave them the news they so badly needed to hear: ‘‘Apollo 8, you are Go for TLI!’’

Drifting high above the Pacific Ocean at the time, the astronauts knew that the burn would be entirely controlled by the computers and, with ten seconds to go, a flashing number ‘99’ appeared on the command module’s display panel. In essence, it asked them to confirm that they wanted to go ahead with the specific manoeuvre. Lovell punched the ‘Proceed’ button and at 10:38 am, some two hours and 47

Had the Saturn V risen or had Florida sunk? On 9 November 1967, the maiden flight of the mighty Saturn V got underway with “naked power, lots of noise and light”. A little more than a year later, Apollo 8 would carry its first human crew.

minutes into the mission, the third stage ignited with a long, slow push. Although Borman kept a keen eye on his instruments in the event that he had to assume manual control, Collins relayed updates from the trajectory specialists that Apollo 8 was in perfect shape. It did not feel that way to Borman, who was convinced from the intense shaking and rattling that he might be forced to abort the burn. Steadily, as Anders watched the third stage’s propellant temperatures and pressures, they turned from ‘Earth-orbiting’ astronauts to ‘Moon-bound’ adventurers. By the time the S-IVB finally shut down after five minutes and 18 seconds, their velocity had increased from 28,100 km/h to 37,300 km/h – the ‘escape velocity’ needed to reach the Moon. Frank Borman, Jim Lovell and Bill Anders were travelling faster than any human beings had ever flown before.

Surprisingly, though, with no outside point of reference, there was not the slightest sense of the tremendous speed at which Apollo 8 was moving. Then, when Borman separated the command and service module from the now-spent S-IVB and manoeuvred around to face the third stage, they saw for the first time the effect of TLI: their home world, Earth, was no longer a seemingly-flat expanse of land and sea and cloud ‘below’ them, but a planet, spherical, its curvature obvious in the black void. They could actually see it receding from them as they continued travelling outwards. At length, as their altitude increased, Earth grew so small that it seemed to fit neatly inside the frame of one of the command module’s windows, then could be easily hidden behind a thumb. ‘‘Tell Conrad he lost his record,’’ Borman radioed Collins. Jim Lovell promptly launched into a geography lesson and even asked Collins to warn the people of Tierra del Fuego to put on their raincoats, as a storm seemed to be approaching.

Manoeuvring Apollo 8 with its nose pointed toward Earth and the S-IVB had not been done simply for sightseeing: Borman’s next task was to rendezvous with it, just as future crews would need to do in order to extract their lunar modules from the enormous ‘garage’ atop the S-IVB. After completing this demonstration, he pulled away for the final time and Apollo 8 set sail for the Moon. Five hours into the flight, after finally removing his space suit, Lovell set to work taking star sightings with the 28-power sextant and navigation telescopes. If they lost contact with Earth, he might have to measure the angles between target stars and the home planet and punch the data into the computer to figure out their position. He would do the same in lunar orbit, measuring craters and landmarks to help refine Apollo 8’s flight path. Not for nothing was ‘navigational expert’ one of the senior pilot’s main responsibilities.

Shortly after 6:00 pm, the first test firing of the SPS engine was performed, lasting just two seconds, which satisfied the astronauts and ground controllers that it could operate as advertised. As the first workday of Apollo 8 drew to a close, Lovell and Anders watched the instruments whilst Borman, unsuccessfully, tried to sleep.

Heading across the vast cislunar gulf, more than 370,000 km wide, the astronauts awakened the first sensations of space sickness. Borman, it seemed, suffered the most. A number of cases of gastroenteritis had plagued Cape Kennedy in the days before launch and it was suggested that this ‘24-hour intestinal flu’ could have triggered the malady; alternatively, Borman had taken a Seconal tablet to help him sleep and blamed the medication for his discomfort. Upon awakening to begin his

The S-IVB recedes into the blackness of cislunar space as Apollo 8 heads for the Moon.

second day aloft, he suffered both vomiting and diarrhoea, but recovered sufficiently by the third day to tell Mission Control that “nobody is sick”. Unknown to Borman, his “case of the 24-hour flu” had caused much consternation amongst the flight surgeons on the ground and even led to suggestions that the mission might have to be terminated. Fortunately, all three men were indeed fine and, even if they were ill, the SPS could not be fired to about-face them back to Earth. They were heading for the Moon, whether they liked it or not.

Strangely, since the Moon was barely a crescent to them at the time, none of the crew really saw it until shortly before their arrival. “I saw it several times in the optics as I was doing some sightings,” admitted Lovell, but “by and large, the body that we were rendezvousing with – that was coming from one direction as we were going to another – we never saw … and we took it on faith that the Moon would be there, which says quite a bit for ground control.” As they headed towards their target, Apollo 8 slowly rotated on its axis in a so-called ‘barbecue roll’, to even out thermal extremes of blistering heat and frigid cold across its metallic surfaces.

Two hundred and twenty-three thousand kilometres from Earth, approximately two-thirds of the way to the Moon and 31 hours since launch, they began their first live telecast from Apollo 8. Borman had tried to have the camera removed from the mission, but had been overruled, and now found himself using it to film Jim Lovell in the command module’s lower equipment bay, readying a dessert of chocolate pudding. Next there was a shot of Bill Anders, twirling his weightless toothbrush. ‘‘This transmission,” Borman commenced for his terrestrial audience, ‘‘is coming to you approximately halfway between the Moon and the Earth. We have about less than 40 hours to go to the Moon… I certainly wish we could show you the Earth. Very, very beautiful.’’

Unfortunately, a telephoto lens fitted to the camera by Anders did not work and when they switched back to the interior lens it resolved the home planet as little more than a white blob, giving away little of its splendour. Borman was disappointed that he had been unable to show viewers the ‘‘beautiful, beautiful view, with blue background and just huge covers of white clouds’’. Lovell closed out the transmission by wishing his mother a happy birthday, after which Borman placed Apollo 8 back into its barbecue roll, which took the high-gain antenna off Earth. A day later, their second telecast was somewhat better, allowing Lovell to describe for his spellbound audience the appearance of the western hemisphere: the royal blues of the deep ocean trenches, the varying browns of the landmasses, the bright whites of the cloud structures.

Lovell was an explorer at heart. His excitement in wanting to fly Apollo 8 was motivated equally as much, if not more so, by the simple urge to explore and see new sights and places than by a desire to carry out scientific investigations. The science was important, but Lovell’s sentiment could perhaps be best tied to a statement made three years later by Apollo 15 commander Dave Scott: that going to the Moon was “exploration at its greatest’’. At one stage in the flight, Lovell turned to Borman and wondered aloud what alien travellers might think as they approached Earth. Would they believe it to be inhabited or not? Would they decide to land on the blue or the brown part of its surface?

“You better hope that we land on the blue part,” deadpanned Anders.

By the afternoon of 23 December, almost 60 hours since their Saturn V left Earth, the gravitational influence of their home planet was finally overcome by that of the Moon. At this point, Apollo 8 was more than 300,000 km from Earth and just 62,600 km from its target and the spacecraft’s velocity had slowed to 4,320 km/h as it moved farther into its gravitational ‘well’. As they sailed towards Lunar Orbit Insertion (LOI), their trajectory was near-perfect: only two of four planned mid-course correction burns had been needed to keep Apollo 8 locked into its free return trajectory. At 3:55 am on Christmas Eve, Capcom Gerry Carr, a member of the Apollo 8 support crew, radioed Borman with the news that they were ‘‘go for LOI’’.

The three astronauts had still not seen the Moon, despite their close proximity to it, since their angle of approach caused it to be lost in the Sun’s glare. At length, Carr asked them what they could see. ‘‘Nothing,’’ replied Anders gloomily, adding ‘‘it’s like being on the inside of a submarine’’. Less than an hour later, at 4:49 am, Apollo 8 passed behind the Moon, with Lovell telling Carr that ‘‘we’ll see you on the other side’’. Eleven minutes later, moving at 9,300 km/h and ‘backwards’, they fired the SPS engine for four minutes to reduce their speed by 3,200 km/h and brake themselves into a 111 x 312 km orbit. The burn was flawless, although Lovell admitted that it was ‘‘the longest four minutes I ever spent’’. Had the engine burned too long or too short, they could have ended up either crashing into the Moon or vanishing into some errant orbit. Just to be sure, Borman hit the shutdown button as soon as the clock touched zero.

Back on Earth, a tense world – nearly a billion people were listening in, NASA estimated, scattered across 64 different countries – waited for word of their insertion into lunar orbit. If Apollo 8 had not fired the SPS, then Borman, Lovell and Anders would come back into communications range ten minutes sooner than planned. At length, right on time, following a 45-minute blackout, public affairs officer Paul Haney announced with joy: ‘‘We got it! We got it!’’ Fifteen minutes later, the astronauts’ first close-range descriptions of the Moon came across more than three hundred thousand kilometres of emptiness: ‘‘The Moon,’’ Lovell began, ‘‘is essentially grey; no colour; looks like plaster of Paris or sort of a greyish deep sand. We can see quite a bit of detail. The Sea of Fertility doesn’t stand out as well here as it does back on Earth. There’s not as much contrast between that and the surrounding craters. The craters are all rounded off. There’s quite a few of them; some of them are newer. Many of them – especially the round ones – look like hits by meteorites or projectiles of some sort… ’’

The lack of even the slightest vestiges of an atmosphere lent a weird clarity to what was, in effect, a scene of the utmost desolation, silence and stillness; the Moon was literally a world frozen in time. Only weeks earlier, the film of Arthur C. Clarke’s ‘2001: A Space Odyssey’ had premiered and even the astronauts imagined the lunar terrain to be composed of dramatic, sharp-edged mountains and jagged cliffs. Instead they were presented with an essentially dead place, seemingly ubiquitous in its dullness and blandness. Anders, tasked with the bulk of the lunar photography, had spent hours before launch with the only geologist-astronaut, Jack Schmitt, discussing the features of the surface, and had his own flight plan to plough

The lunar farside, never before seen directly by human eyes.

through, but found it hard because of dirty windows. In fact, only the command module’s two small rendezvous windows remained reasonably clear.

For Anders, the far side of the Moon, never seen from Earth or ever by human eyes, resembled “a sand pile my kids have been playing in for a long time… it’s all beat up, no definition, just a lot of bumps and holes’’. He considered the lunar surface to be an unappealing place, albeit with “a kind of stark beauty’’ of its own, and all three men found pleasure in giving temporary names to some of the craters to honour their colleagues and managers: Low, Gilruth, Shea, Grissom, White, Webb, Chaffee, Kraft, See, Bassett and others. “These,” said Borman, “were all the giants who made it work.’’ At one stage during the excitement, when flight controller John Aaron noticed that the command module’s environmental control system needed adjustment, they responded by naming a crater for him, too. (Before the flight, Lovell had even given his wife Marilyn a photograph of a mountain, near the eastern edge of the Sea of Tranquility, which he had unofficially named for her: Mount Marilyn.)

Four hours after entering orbit, another SPS burn, this time thankfully shorter at just 11 seconds, adjusted Apollo 8’s path around the Moon into a near-perfect 111 km circle. Then, at 10:37 am on 24 December, their first glimpse of colour entered a Universe of endless blackness and greyness: the three astronauts became the first humans to witness ‘Earthrise’ from behind the lunar limb. Borman was in the process of turning the spacecraft to permit Lovell to take some sextant readings, when all at once Anders yelled: ‘‘Oh my God! Look at that picture over there.’’ It would become a running, though light-hearted competition among the crew over who took the ‘Earthrise Picture’ which has since become world-famous: a shot of our home planet, a pretty blue-and-white marble, rising in the void above the Moon’s grey-brown surface. With Lovell in attendance, it was Anders who, after fitting the colour magazine and aiming the telephoto lens, snapped one of the most iconic images of the Space Age. In perhaps no other image has the beauty, fragility and loneliness of Earth been captured with more meaning. Years later, Anders would win praise from environmentalists for his assertion that Apollo 8’s goal was to explore the Moon. . . and what it really did was discover the Earth!

The astronauts’ intense workload during their 20 hours in orbit was getting the better of them, with tiredness causing them to make mistakes. On occasion, Lovell had punched the wrong code into the command module’s computer, triggering warning alarms, and Anders was overcome with his own schedule: stereo imagery, dim-light photography and filter work. At length, clearly irritated that the timeline was too full, Borman snapped at Capcom Mike Collins that he was taking an executive decision for his two crewmates to get some rest. ‘‘I’ll stay up and keep the spacecraft vertical,’’ he told Collins, ‘‘and take some automatic pictures.’’ With some difficulty, he had to force Lovell and Anders to pry their eyes away from the windows and get some sleep.

It seemed inevitable, after thousands of years of watching and wondering about the Moon, that humanity’s first visit would be commemorated in a religious, spiritual or symbolic way. Before the launch, Borman, Lovell and Anders had discussed this issue at length with friends and concluded that they would read the story of Creation from the first ten verses of Genesis. During their ninth orbit, on their second live telecast from the Moon, they read it to a spellbound world, first Anders taking a part, then Lovell and finally Borman closing with ‘‘Good night, good luck, a Merry Christmas and God bless all of you. . . all of you on the good Earth’’.

Eight minutes into Christmas morning, three days and 17 hours after launch, the return home got underway when the SPS engine was ignited to increase their speed by 3,800 km/h. As they rounded the Moon for the last time, Lovell told Capcom Ken Mattingly, who was just coming on duty in Houston, ‘‘Please be informed there is a Santa Claus’’. Mattingly replied that they were the best ones to know.

The return journey proved uneventful, with fogged windows, puddling water and clattering cabin fans creating mere annoyances. A final televised tour of Apollo 8 showed Anders preparing a freeze-dried meal. . . and, when the camera stopped rolling, they found a real treat in their food locker: real turkey and real cranberry sauce, wrapped in foil with red and green ribbons. It was a far cry from the

One of the 20th century’s most iconic images: Earthrise from Apollo 8.

toothpaste tubes of Project Mercury and even better, perhaps, than Gus Grissom’s corned beef sandwich. The turkey and cranberry sauce turned out to be their best meal of the entire flight, although Borman was annoyed that Deke Slayton had slipped three small bottles of brandy aboard as well. Why, if anything went wrong on the flight, the overzealous Borman fumed, the press and public would blame it on the ‘drunk’ astronauts. Lovell and Anders, who have admitted that they had no intention of touching the brandy, felt that Borman had gone a little too far. Christmas spirit returned, however, with festive presents: pairs of cufflinks and a man-in-the-Moon tie pin from Susan Borman and Marilyn Lovell and a gold ‘figure 8’ tie pin from Valerie Anders.

Only one minor trajectory correction burn was needed and early on 27 December, the astronauts fired pyrotechnics to jettison the service module and plunged into Earth’s atmosphere at 34,900 km/h. During re-entry, which carried them over north­eastern China, then brought the command module in a long slanting path towards the south-east, Borman, Lovell and Anders were subjected to deceleration forces as high as 7 G. Splashdown came as Cape Kennedy clocks read 10:51 am, but still in pre-dawn darkness over the western Pacific, completing a mission of just over six days. At Mission Control in Houston, sheer pandemonium broke out, in the traditional American back-slapping way, and the smell of celebratory cigars scented the air for hours.

Among the cheering NASA throng was an overjoyed, though dejected Mike Collins. ‘‘For me personally, the moment was a conglomeration of emotions and memories,’’ he wrote. ‘‘I was a basket case, emotionally wrung out. I had seen this flight evolve in the white room at Downey, in the interminable series of meetings at Houston… into an epic voyage. I had helped it grow. I had two years invested in it – it was my flight. Yet it was not my flight; I was but one of a hundred packed into a noisy room.’’

A quarter of a world away, in the Pacific Ocean, some 1,600 km south-south-west of Hawaii, water came flooding through an open vent in the command module, drenching Borman and giving Anders the mistaken impression that the hull had cracked on impact. The ship overturned onto its nose, but quickly righted itself when Borman inflated the three airbags. It did not stop him from being sick. This time, Lovell and Anders, both of whom had served in the Navy, showed no mercy on their Air Force commander: ‘‘What do you expect from a West Point ground-pounder?’’

Amidst the radio chatter from a rescue helicopter despatched by the aircraft carrier Yorktown came an age-old question which the whole world now wanted answered. ‘‘Apollo 8, is the Moon made from Limburger cheese?’’

‘‘Nope,’’ replied Bill Anders. ‘‘It’s made from American cheese!’’

TRAINING

Of course, wrote Scott, he had to be sure that Armstrong, as Gemini VIII’s command pilot, was happy with his selection. He need not have worried. “When I caught up with him in a corridor outside the VIP room,’’ Scott related, “he gave me a big grin and held out his hand. That was all I needed.’’

The official announcement came on 20 September, with Pete Conrad and Dick Gordon assigned as their backups. Gemini VIII would be the most complex mission ever attempted, involving all of the tasks practiced by previous flights… and more: three days in orbit, rendezvous – and, for the first time, physical docking – with an Agena-D target, combined manoeuvres, scientific experiments, a precision re-entry and a tricky spacewalk. The sheer complexity of the flight, wrote Scott, “was reflected in our mission badge. . . a rainbow of colours refracted from twin white stars through a prism to form the Gemini symbol together with a Roman ‘VIII’ … to reflect its many objectives’’.

Scott’s EVA was part of a goal that simply had to be perfected before the early Apollo missions and the Department of Defense, through the Air Force, had already invested heavily in an Astronaut Manoeuvring Unit (AMU) to be tested by Charlie Bassett on Gemini IX. Before that, Dave Scott would wear a suit with a chest pack (known as the Extravehicular Life Support System) to feed oxygen from the spacecraft’s supply and from a backpack (dubbed the Extravehicular Support Package) located on the spacecraft’s rear adaptor section. This backpack also provided a radio and 8 kg of propellant for a zip-gun manoeuvring tool – similar to the hand-held device used by Ed White – and was connected to Gemini VIII’s systems by an 8 m oxygen-hose tether.

When Scott reached the rear of the spacecraft and backed himself into the backpack, he would add a lightweight 23 m tether to the hose, allowing him to move up to 30 m into space. The exercise was extremely risky and if Scott encountered problems, there would be no way for Armstrong to aid him: the command pilot therefore requested a realistic model of the Gemini VIII adaptor for training, to rehearse donning the backpack, together with practice runs in an altitude chamber. Scott, meanwhile, worked closely with Ed White and by the eve of launch had completed more than 300 zero-gravity aircraft parabolas and over 20 hours on an air-bearing table, donning and doffing the bulky suit again and again. It was hot, hard, strenuous work. ‘‘I remember Dick [Gordon] turning to me once,’’ wrote Scott, ‘‘drenched in sweat and joking ‘Isn’t this glamorous?’’’

Conversations with White had certainly identified the need for Scott to maintain

TRAINING

The space suit and equipment scheduled to be used by Dave Scott on Gemini VIII.

his physical fitness and strength, particularly in light of the difficulties in closing the hatch on Gemini IV. A consequence of this was that a lever was added to Gemini VUI’s hatch, allowing it to be closed with less physical force. Nonetheless, White cautioned, the space suit itself was both stiff and heavy, requiring strength and stamina to move around for two hours. Scott began jogging, playing handball and lifting weights in the gym. . . and, in his autobiography, would recall how this brought him face to face with Neil Armstrong’s sly humour.

Armstrong’s theory on exercise, John Glenn once said, was that a human being had only a finite number of heartbeats and should not waste them frivolously. One day during Gemini VIII training, as Scott pumped iron in the gym, Armstrong turned up, set the exercise bicycle onto its lowest possible tension setting and began pedalling, telling him: “Attaboy, Dave! Way to go!’’

Work on the 6 x 7 m air-bearing table – which pumped highly-compressed air through holes in its floor to remove friction – allowed Scott to literally ‘fly’ across its surface using the zip-gun, which he described as one of the most valuable parts of his EVA training. The gun had 15 times more propellant than White’s device and, instead of oxygen, used freon, a refrigerant with a correspondingly higher density. This multiplied the gun’s impulse, although Scott worried how the gas might behave in space. During one test at low temperatures, for example, the freon caused the gun’s poppet valve to stick open when triggered. Had this occurred in space, the escaping gas might have caused Scott to tumble uncontrollably. New seals were fitted to solve the problem and two new shut-off valves provided additional safety.

The space suit’s other equipment presented its own problems. One key obstacle was the risk that an ejector in the chest pack could freeze and block the flow of oxygen from both Gemini VIII’s supply and from the backpack. In response, engineers incorporated 20-watt heaters near the ejector. Other obstacles included overcoming the jumble of umbilicals, tethers and jumper cables whilst donning the chest pack inside the cramped cabin, although by December 1965 Scott was satisfied that he could do the task satisfactorily. The explosion of Wally Schirra and Tom Stafford’s Agena-D, on the other hand, did not fill Armstrong or Scott with confidence. Mission Control shared their concerns, telling them to ‘‘get out fast’’ if the target showed the slightest hint of a malfunction whilst docked.

Despite the EVA experience gained by Ed White, the Gemini VIII excursion would eclipse it in complexity. Not only would Scott be outside for two hours and 40 minutes – nearly ten times longer than White – but the flight plan also required him to move to the rear of the spacecraft to don his backpack. He would spacewalk through orbital daytime and nighttime, retrieve an emulsion package, activate a micrometeoroid collector on the Agena and test a reactionless power wrench. As a result, in addition to aircraft parabolas and trial runs on the air-bearing table, he was submerged into a large water tank with a mockup zip-gun to practice moving around in a neutrally-buoyant environment. Today, such underwater exercises are common as spacewalkers rehearse procedures outside the International Space Station and are regarded by many as the closest analogue to the real thing.

Training for the spacewalk was, however, only part of the Gemini VIII mission; Armstrong and Scott’s 70 hours aloft also featured a full plate of experiments. One

TRAINING

Artist’s concept of Dave Scott during his EVA.

of these was the Zodiacal Light Photography investigation, which sought to capture a series of electronically-controlled exposures using a modified 35 mm Widelux camera with a rotating lens. “This was to observe… a very faint glow on the horizon on Earth seen just after sunset and before sunrise,” Scott wrote. To accustom their eyes to the subtle zodiacal light, caused by dust circling towards the inner Solar System, Armstrong and Scott would turn down the lights in their T-38 cockpit during cross-country night flights.

Also aboard Gemini VIII was a hand-held spectrometer to deduce cloud-top heights from the absorption of light by the oxygen band, together with a nuclear emulsion experiment to determine cosmic radiation effects on bromide crystals, a micrometeoroid investigation affixed to the Agena and a series of synoptic terrain photography tasks for meteorological research.

THE END

As the Gemini XII flight hardware – Lovell and Aldrin’s spacecraft home, their Titan booster, their Atlas-Agena – was readied for launch during 1966, it was accompanied by an impending deadline to terminate the project and press on with Apollo. Indeed, when the two men walked out to Pad 19 on launch morning, 11 November, they would wear placards reading ‘The’ and ‘End’ on their backs. By this point, the corroded dome of the second-stage fuel tank previously assigned to Gemini X had been repaired and was delivered from Martin’s Denver to Baltimore plants in late January. Seven months later, on 12 August, the entire Titan for Gemini XII was approved by Martin and by mid-September both of its stages were at Cape Kennedy.

It was at around this time that the Air Force’s AMU test was deleted from Gemini XII and Gene Cernan lost his final opportunity to test-fly it in orbit. Persistent problems with mastering EVA techniques on previous missions had, NASA management concluded, made it inadvisable to proceed with such an ambitious endeavour and Gemini XII would instead focus on perfecting the ‘fundamentals’ with just basic extravehicular tasks. During his time outside, Aldrin would remove, install and tighten bolts with the power tool whose evaluation had been denied both Dave Scott and Dick Gordon, as well as operating connectors and hooks, stripping patches of Velcro and cutting cables.

To physically condition themselves, Aldrin and Cernan spent a considerable amount of time underwater in the neutral buoyancy tank, just outside Baltimore. They wore carefully-ballasted suits, Aldrin said later, to completely neutralise their buoyancy and approximate microgravity conditions as closely as possible. “Eventually,” he wrote, “I mastered the intricate ballet of weightlessness. Your body simply had to be anchored, because if it wasn’t, flexing your pinkie would send you ass-over-teakettle. And you don’t want to do that dangling at the end of an umbilical cord 160 miles above Earth.’’

Processing of the Atlas booster and Gemini’s final Agena target ran in tandem with that of the Titan. Designated Agena ‘5001’, it was actually the non-flying version of the target delivered to Cape Kennedy in the summer of 1965, which had since been upgraded and made space-capable. ‘‘Getting a docking target took a bit of juggling,’’ wrote Deke Slayton. The Air Force formally accepted it for advanced processing early in September 1966 and by the end of October it had been mated atop its Atlas and installed on Pad 14. For Lovell and Aldrin, their scheduled launch just a few days after Halloween had spawned an interesting orange-and-black embroidered crew patch, together with a crescent Moon offering a nod to the impending Apollo project.

Plans to launch on 9 November were abandoned when a malfunctioning power supply in the Titan’s secondary autopilot reared its head and Lovell and Aldrin were recycled to fly two days later. The morning of the 11th dawned fine and clear and the Agena set off promptly that afternoon at 2:08 pm. (During insertion into space, an anomaly was noted in the target’s propulsion system and plans to boost Gemini XII into a higher orbit were abandoned.) Strapped inside their tiny cabin, both astronauts could clearly hear the Atlas’ thunderous roar. Ninety-eight minutes later, at 3:46:33 pm, it was their turn.

‘‘There was no noise at first,’’ Aldrin wrote, ‘‘but then a growing rumble began as the spacecraft rolled through its pre-programmed manoeuvre, twisting to the proper south-east launch trajectory.’’ Steadily, the Titan accelerated, ‘‘like a subway train’’, Aldrin recalled, and as they climbed ever higher the sky turned to dark blue and eventually to black. Inside their space suits, both men felt their limbs rise and their toes lift to touch the tops of their boots. It felt almost as if they were stretching their feet, but not quite. They were weightless.

Once established in their 160 x 270 km orbit, Lovell and Aldrin set to work ploughing through their checklists, preparing for rendezvous and docking with the Agena some three orbits – and a little over four hours – into the mission. At around 5:11 pm, they made their first attempt at radar contact with the target and were surprised when the computer responded with the desired digits. ‘‘Houston,’’ radioed a jubilant Aldrin, ‘‘be advised we have a solid lock-on … two hundred thirty-five point fifty nautical miles.’’

However, the astronauts’ success proved short-lived. As they circularised their

THE END

Jim Lovell (left) and Buzz Aldrin at breakfast on launch morning.

 

Подпись: The end 387

orbit to align themselves ‘behind’ and ‘below’ the Agena, above North America, Gemini XII’s radar began giving intermittent readings. It was at this stage that Aldrin’s years of rendezvous work came to the fore: he broke out the intricate charts and reverted to what he called the ‘Mark One Cranium Computer’ – the human brain. In his autobiography ‘Men from Earth’, Aldrin vividly described the hours – long effort: as Lovell piloted Gemini XII, he laboured over the charts, barely able to see the closely-printed data, occasionally aware of the passage of orbital daytime into nighttime and vice-versa.

It paid off. A little over four hours into the mission, Lovell eased the spacecraft’s nose into the Agena’s docking collar and announced, somewhat nonchalantly, ‘‘Houston, we are docked’’. The response from the ground, delivered with similar excitement, was a simple ‘‘Roger’’. A potentially serious obstacle – the failure of a critical piece of equipment, the rendezvous radar – had been overcome by human brainpower and flying abilities. Should a similar contingency occur during a rendezvous situation in orbit around the Moon, Lovell and Aldrin’s work had at least proved that workarounds could be achieved. They had also used barely 127 kg of their fuel supply in one of the project’s most economical rendezvous.

For the fourth time in eight months, a Gemini was securely linked to an Agena and Lovell and Aldrin became the second crew to practice undocking and redocking exercises. One attempt by Lovell during orbital darkness caused the docking latches to ‘hang up’ – producing a rather disturbing grinding sound – but he was nevertheless able to rock Gemini XII free without damage. A few minutes later, they switched roles and Aldrin redocked them onto the target.

Original plans, laid out before launch, had called for a reboost to high altitude, but this had to be abandoned eight minutes after the Agena lifted-off when its engine suffered a momentary decay in thrust chamber pressures and a drop in turbine speed. Instead, the astronauts were directed to turn their attention to solar eclipse photography; this task had been a scheduled part of their mission had they launched on 9 November, but the two-day delay caused it to be dropped. Now that the Agena reboost had been cancelled, it was reinstated, thanks to the input of Gemini XII’s experiments advisory officer James Bates.

The inclusion of Bates’ recommendation marked a shift in operations, with the scientists’ representative, for the first time, being allowed to participate as one of the flight control team in the main Mission Control room. Moreover, it was determined that the Agena’s secondary propulsion system had enough power to orient the spacecraft for an eight-second photographic pass at the proper time. At 10:51 pm, a little over seven hours into the mission, Lovell duly fired the target’s smaller engines to reduce the combination’s speed by 13 m/sec. The adjustment was successful and, after their first sleep period, the astronauts were advised to perform a second firing. Sixteen hours after launch, they reported seeing the eclipse ‘‘right on the money’’, cutting a swath across South America from north of Lima down to the southernmost tip of Brazil.

At first, it had seemed to the disgruntled crew that the second Agena burn might throw out the remainder of their schedule and adversely affect the start of Aldrin’s first EVA. It did not, and at 11:15 am on 12 November, some 20 minutes before orbital sunset, Aldrin cranked open his hatch and pushed his helmeted head outside. “The hatch rose easily,” he wrote, “and I rose with it, floating above my seat, secured to the spacecraft by short oxygen inflow and outflow umbilical hoses.” Years later, he would vividly describe the immensity of the Universe all around him, remember the absence of any sense of speed and the recall the distinct curvature of Earth.

Aldrin quickly set to work on his first task, dumping a small bag containing used food pouches, and watched it slowly tumble away like a top, straight ‘above’ him. Next he moved on to quickly attach cameras onto brackets to photograph star fields on ultraviolet film and retrieved a micrometeoroid package, which he passed inside to Lovell. Unlike Cernan and Gordon, he did not overheat, thanks partly to regularly-scheduled rest breaks of two minutes apiece, and he returned inside Gemini XII at 1:44 pm after two and a half hours.

His real work had yet to begin. The mission’s second period of EVA, which got underway at 10:34 am the following morning, required Aldrin to move away from the spacecraft on a 9 m tether. He set up a movie camera to allow flight controllers to monitor his performance, then moved to Gemini XII’s nose and affixed a waist restraint strap to the docking adaptor. Next, Aldrin removed a tether from the Agena’s nose and snapped it onto the Gemini, connecting the two vehicles for a gravity gradient exercise scheduled for later in the mission. He then manoeuvred himself towards the rear of the spacecraft, using flatiron-shaped handholds fitted with Velcro patches, and slipped his boots into a pair of foot restraints nicknamed ‘golden slippers’. These, coupled with two small waist tethers, kept him anchored securely and Aldrin was able to satisfactorily complete a number of tool-handling and dexterity tests.

‘‘Back in the buoyancy pool in Maryland,’’ he wrote later, ‘‘I had torqued bolts and cut metal dozens of times – what I used to call ‘chimpanzee work’ – and I had no problem with these chores in space. Someone even put a bright yellow paper Chiquita Banana sticker at my busy box.’’ He was even able to wipe Lovell’s window (who asked him to change the oil, too) before returning to the cabin after two hours and six minutes outside. Back on Earth, Aldrin would claim quite openly that he had personally solved many of the problems of EVA, arousing criticism among the other astronauts, including Gene Cernan, who felt that his tasks were nowhere near as difficult as theirs. ‘‘Quite frankly,’’ Cernan wrote, ‘‘we said he was only working a monkey board. Draw your own conclusions.”

Shortly after Aldrin’s return inside Gemini XII, the two men completed their evaluation of the tether by undocking from the Agena. The tether tended to remain slack, although they believed that slow gravity gradient stabilisation was achieved. ‘‘Within minutes,’’ wrote Aldrin, ‘‘the two vehicles had stabilised without the aid of thrusters.’’ After two full orbits thus connected, they finally fired an explosive squib to jettison the tether at 7:37 pm on 13 November.

Aldrin’s record-breaking five and a half hours of cumulative extravehicular experience concluded the following day, the 14th, when he ventured outside at 9:52 am for a second stand-up period, lasting 55 minutes. He dumped unneeded equipment overboard, together with a sack containing his umbilical tether and two rubbish bags – hurled in lazy arcs over his shoulder – and took one last lingering

THE END

Aldrin during one of his three periods of EVA.

look at Earth below him: the vast land mass of Indochina… and thought of his friend, Sam Johnson, with whom he had undergone flight training, and who was at that very moment a prisoner of war somewhere in North Vietnam.

Lovell and Aldrin’s four-day mission had brought Project Gemini to a spectacular conclusion and had satisfactorily demonstrated rendezvous, docking, gravity gradient tethered operations and the ability of skilled human pilots to calculate a rendezvous with sextants and charts and a slide rule and pencil. Such human skills, using, in Aldrin’s own words, the Mark One Cranium Computer, had relaxed managers’ concerns about the viability of astronauts being able to perform a manual rendezvous, if necessary, in orbit around the Moon.

Gemini XII’s problems were comparatively minor. Four of its 16 thrusters failed during the course of the mission and two its six fuel cells went dead, obliging flight controllers to instruct Lovell and Aldrin to drink more than their planned rations of water. This would make room for the excess fuel-cell water, which otherwise threatened to flood the spacecraft’s power system. Whenever they drank water or used it to prepare their food, the red warning light blinked off, and in this way they nursed the fuel cell through 80 hours of flight.

A re-entry controlled completely by the computers brought Gemini XII into the Atlantic, barely 4.8 km from its target impact point, at 2:21 pm on 15 November. Within half an hour of splashdown, Lovell and Aldrin were safely aboard the aircraft carrier Wasp. The only unexpected event during re-entry had come at the onset of peak G loads, when a pouch containing books, filters and equipment broke free from the sidewall and landed on Lovell’s lap. By this time, both men had unstowed the D-rings for their ejection seats and Lovell fought the urge to catch the pouch, lest he accidentally grab and pull the ring. “I didn’t want to see myself punching out right at this high heating area,’’ he said later.

With the safe return of Lovell and Aldrin to Earth, many of the procedures needed to get to the Moon and back had been thoroughly tested. Extravehicular suits had been used for extended periods of time and five astronauts had completed useful tasks outside. Unlike Alexei Leonov’s swim in the void 20 months earlier, they had actually begun to demonstrate an astronaut’s ability to really work in space. It provided the closest analogue yet attained of what working on the lunar surface might be like. Rendezvous, despite its complexity, had been completed with seemingly effortless ease by six Gemini command pilots… and Lovell and Aldrin’s work had shown it could be done without the aid of radar.

The radiant Moon above Cape Kennedy in the early winter of 1966-67 seemed considerably brighter than normal, as an altogether different kind of space vehicle geared up for its first manned shakedown cruise. Sitting on Pad 34 was a far larger rocket – the Saturn 1B – topped with the Apollo 1 spacecraft. In February 1967, astronauts Gus Grissom, Ed White and Roger Chaffee would evaluate the machine that would carry Americans to the Moon. Forget 1970, said many within NASA; it was becoming increasingly likely that a lunar landing might be achieved two years ahead of schedule. For ten euphoric weeks from mid-November 1966, the Moon was within humanity’s grasp. Then, on the fateful Friday evening of 27 January 1967, all such dreams dissolved.

6

SAVED

A few weeks after Apollo 8 splashed down, amongst his mountains of fan mail, Frank Borman came across a telegram from a stranger which summed up the entire mission and the effect of the mission in three words. It read simply: ‘‘You saved 1968’’.

In spite of Borman, Lovell and Anders’ achievement, the telegram’s sender was right in that the year had been a bad one, both in America and elsewhere. Israel and Palestine clashed in border disputes, three decades of ‘Troubles’ began in Northern Ireland, Soviet tanks rolled in Czechoslovakia to stifle the Prague Spring reforms, the Vietnam War seemed unwinnable and so unpopular that President Lyndon Johnson had glumly announced in March that he had no intention of running for re­election and the United States was left reeling by the murders of Martin Luther King and Senator Bobby Kennedy. Apollo 8, humanity’s first journey to the Moon, had cast one of few rays of light over a desperately unhappy and violent year.

On 29 March, King had visited Memphis, Tennessee, in support of black sanitary works employees, who were striking for higher wages and better treatment. A few days later, he delivered his famous ‘I’ve Been To The Mountaintop’ speech, then checked into his room at the Lorraine Motel. At precisely 6:01 pm on 4 April, as he stood at his balcony, he was shot; the bullet passing through his right cheek, smashing his jaw, travelling down his spine and finally lodging somewhere in his shoulder. In spite of emergency surgery, the man who had fought tirelessly for civil rights in America was dead. . . and his murder instantly sparked fury in as many as a hundred cities across the nation. Two months later, on 5 June, Senator Kennedy – brother of the murdered president, one-time attorney-general and having himself just won the California primary as part of his own presidential candidacy bid – was shot in the crowded kitchen passageway of the Ambassador Hotel in Los Angeles. Both killings provoked desperate outrage as two men who supported two of the strongest issues of the day – civil rights and ending the Vietnam War – were prematurely cut down.

As the last few days of blood-stained 1968 faded into history, however, the long road to the Moon had been won. The enormous technological challenges needed to navigate men and machines across a gulf of more than three hundred thousand kilometres of uncharted emptiness had been met. Nor was Apollo 8 simply a lucky shot: Frank Borman, Jim Lovell and Bill Anders had been guided into a precise orbit around our closest celestial neighbour and had taken a truly giant leap towards the small step which, seven months hence, would change humanity’s view of itself forever. Completion of the first circumlunar mission was just the first part of John Kennedy’s promise. Now, in the final few months of the decade, would come the most audacious task of all: achieving all that Apollo 8 had achieved and more, guiding the spidery, as-yet-untested lunar module down those last 111 km from orbit and planting American bootprints onto the Moon’s dusty surface. The pieces were set. The machines, equipment and rockets were ready. So were the men. On 6 January 1969, Deke Slayton summoned Neil Armstrong, Mike Collins and Buzz Aldrin into his office. In just two words, he told them the news every astronaut had trained for years to hear: ‘‘You’re it!’’

‘It’, of course, meant that they were being tapped for the coveted lunar landing, tentatively pencilled-in for Apollo 11 in July. Armstrong and Aldrin had just come off their Apollo 8 backup duties and Slayton, perhaps, felt a pang of conscience for Collins, who had fought his way to full health and back onto flight status after his neck surgery. There would be a slight shift of roles, though. Following Jim Lovell’s departure for the Borman crew, Aldrin had been promoted to senior pilot of the Apollo 8 backup team. It was a role later to become synonymous with the command module pilot, essentially a mission’s second-in-command, but both Armstrong and Slayton had more confidence in Collins to fill this role. “I had a little difficulty putting Aldrin above Collins,’’ Armstrong told James Hansen. “In talking with Deke, we decided, because the CMP had such significant responsibilities for flying the command module solo and being able to do rendezvous by himself and so forth, that Mike was best to be in that position.’’ Thus, Aldrin missed out on being Apollo ll’s second-in-command, but the alternative was far sweeter: if the schedule ran as planned, he would be the second man on the Moon.

Of course, we know today that Kennedy’s goal was indeed met. Yet a lunar landing on Apollo ll was by no means set in stone as 1969 dawned. (Mike Collins would later estimate the chance of success, in his mind at least, as no more than 50­50.) Still to be proven was Grumman’s spidery lunar module, which Apollo 9 astronauts Jim McDivitt and Rusty Schweickart intended to test in Earth orbit late in February, as their colleague Dave Scott practiced rendezvous and docking in the command module. The space suit which astronauts would one day use to walk on the Moon would be put through its paces by Schweickart during a dramatic EVA, in which he would climb out of the lunar module’s hatch and onto its porch. A couple of months later, Apollo l0 would do a full dress-rehearsal of the Moon landing. . . 370,000 km away, in lunar orbit. Tom Stafford and Gene Cernan, old buddies from Gemini IX-A, would guide the lunar module to just 15 km above the surface, before firing their ascent engine to boost themselves back up to rendezvous with crewmate John Young. Only if both of these highly-complex missions, the details of which remained to be hammered-out, succeeded could Apollo ll stand any chance of launching in mid-July.

The Soviets, too, were on the brink of re-entering the game with a vengeance. More than a year and a half after Vladimir Komarov’s tragic death, the new Soyuz spacecraft was operating and, for them, l969 would see no fewer than five manned missions: two in January which would feature their first spacewalk in almost four years, carried out, finally, by Yevgeni Khrunov and Alexei Yeliseyev, and a unique triple rendezvous involving seven cosmonauts in October, which the Soviets would laud as having laid the foundations of a long-term space station. However, the writing did seem to be on the wall as far as the their chances of getting a cosmonaut onto the lunar surface before the Americans was concerned; Nikolai Kamanin had long since written in his ubiquitous diary that he was convinced the United States would win the race. Yet the lure of the Moon and getting cosmonauts there would not fade from the Soviet psyche for some time and, indeed, l969 would prove a make-or-break year for the enormous, temperamental N-l booster. If that beast – even more powerful than the Saturn V – could somehow be tamed, made to work and entrusted with a human crew, a flag bearing the Hammer and Sickle might still end up sticking out of the lunar soil.

In addition to closing out the first decade of manned spaceflight, l969 offered a starting point for the future: after the G mission, longer stays on the Moon were envisaged, running into the Seventies, with perhaps lunar bases and expeditions to

Mars thereafter. A revolutionary reusable spacecraft known as the Space Shuttle would begin its tumultuous development and it was hoped that, instead of simply visiting the heavens, men would actually come to live there. By the end of the Seventies, as the Shuttle prepared for its maiden launch and promised access to space that was cheaper than ever before, a total of six flags and hundreds of bootprints would dot half a dozen lunar landing sites and Soviet cosmonauts would routinely spend six months at a time in orbit, hosting guests from other nations in their orbiting stations.

Humanity had advanced enormously between the end of the Fifties and the close of the Sixties, in a thousand social, cultural, political and technological ways. It would have been impossible to imagine on the eve of Yuri Gagarin’s orbital flight that within such a short span of time the techniques and tools of rendezvous, docking, spacewalking and reaching the Moon would have been tried, tested and mastered. Four per cent of the federal budget, in Apollo’s case, had much to do with this speed and success, but it remains quite remarkable that the International Space Station has required two decades from conception to construction and at least 15 years will have passed by the time George W. Bush’s vision of humans back on the Moon is realised sometime around 2020.

The Sixties were truly an inspirational, pivotal decade which shaped the future of space exploration. Lessons were learned which have much bearing on activities in orbit today and some continue to be relearned for the missions of the future. Yet they only represented the first few years of a human adventure which, to date, has spanned five decades. If the Sixties involved simply rising from Earth, as Socrates said, and reaching the top of the atmosphere to understand the world from which we came, then the Seventies would establish our first foothold in the heavens.

“LET’S GO!”

His voice exhibited clarity, calmness and confidence when Korolev called over the radio an hour before launch for a status update. To alleviate the boredom, Pavel Popovich, the cosmonaut on duty in the control bunker, arranged for some music to be played. This stopped abruptly at 8:51 am as Korolev gruffly announced that, with a little over 15 minutes to go, it was time for Gagarin to seal his gloves and close his visor. Unlike their counterparts in the United States, no ‘5-4-3-2-1’ countdown was followed; the R-7 was simply fired at the appointed time. As a result, the last few seconds before humanity’s first voyage into space were almost anti-climatic. With steady rhythm, Korolev barked out in turn: ‘‘Launch key to ‘go’ position. . . Air purging… Idle run,’’ and, finally, at 9:07 am, ‘‘Ignition!’’ The gantry’s supporting arms sprang clear from the sides of the rocket as its 20 kerosene-fed engines, with an explosive yield of 880,000 kg, roared to full power.

Gagarin would later describe the initial sensation of liftoff as ‘‘an ever-growing din’’, albeit no louder than the sounds he had experienced flying high-performance jets at Nikel. His helmet muffled much of the R-7’s rumble, although the vibrations remained apparent. At some point, a few seconds after leaving Earth, he yelled the immortal words ‘‘Poyekhali!’’ (‘‘Let’s go!’’) over the radio circuit. Within two minutes, as the G loads began to build, he found it increasingly difficult to speak and would later compare the sensation to the stress of a harsh turn in a MiG. The pressure lifted momentarily as the R-7’s four strap-on boosters burned out and separated; after a brief pause, the central core of the rocket accelerated and the G loads began to rise again. By three minutes into the flight, at 9:10 am, the Little Seven’s nose fairing was jettisoned from around Vostok’s ball, giving Gagarin his first glimpse of the dark blue sky and the clear curvature of Earth as he reached the edge of space.

The R-7’s core, now exhausted of propellant, finally fell away some five minutes into the flight, leaving Gagarin reliant on an upper-stage engine, which inserted him

“LET’S GO!”

Vostok 1 ascends from Tyuratam atop Korolev’s Little Seven. Note the opened petals of the ‘tulip’ launch platform.

 

into orbit promptly at 9:18 am, exactly as Korolev had planned. Unknown to the cosmonaut, the core had actually burned for longer than anticipated, leaving Vostok in an orbit with an apogee – high point – of 327 km, rather than the intended 230 km. It added a little more height to Russia’s already-won World Aviation Altitude Record. As Muscovites arrived at work and the western world still slumbered – figuratively and literally – a new era had begun. Gagarin, whose heart rate soared from 66 to 158 beats per minute during ascent, had won the first lap of the space race.

Although the noise and intense vibration of the R-7 was now gone, it was succeeded by the steady murmur of fans, ventilators, pumps and the hiss of static in his ears. His first experience of the state of weightlessness, properly termed ‘microgravity’, was hampered by the fact that he was tightly strapped into his ejection seat. However, he had purposely carried a small Russian doll as a gravity indicator and watched as it floated comically in midair. So too did a notepad and pencil. In fact, this would not be the doll’s only experience of space travel: in 1991, cosmonaut Musa Manarov would carry it on his mission to commemorate the 30th anniversary of Gagarin’s flight.

As the capsule slowly rotated, to avoid wasting propellant on unnecessary manoeuvres and also to ensure that various sections of Vostok did not become too hot or cold, the First Cosmonaut had his first opportunity to view the world through the Vzor. In his official statement, made in Moscow a few days later, he would describe ‘‘a smooth transition from pale blue, blue, dark blue, violet and absolutely black… a magnificent picture’’. He started jotting down observations in his notepad, but after the pencil floated away he turned instead to the on-board tape recorder to log his thoughts, describing weightlessness as ‘‘not at all unpleasant’’ and confirming that the food and drink were good. Two Russian schoolgirls, who would sample that same food in less than two hours’ time, would be inclined to disagree.

Ten minutes after launch, Vostok was heading eastwards in the direction of Siberia and Gagarin gradually drifted out of radio range with Tyuratam, to be picked up in turn by other listening posts at Novosibirsk, Kolpashevo, Khabarovsk and the easternmost station on Soviet soil, Yelizovo, near Petropavlovsk on the Kamchatka Peninsula. To attract Gagarin’s attention, each station transmitted musical themes to him: Muscovite tunes, ‘Waves of the Amur’, Baglanova songs and others. ‘‘During that early period,’’ Alexei Leonov wrote, ‘‘the Mission Control Centre at Yevpatoriya in the Crimea . . . was still under construction, so communication and control from the ground. . . were performed from the radio stations above which the spacecraft flew.’’ Leonov had been sent to the remote Yelizovo site a few days earlier as its cosmonaut representative and, to underline the secrecy, had no idea if Gagarin or Titov was aboard Vostok. Then, as the spacecraft passed over Kamchatka at around 9:21 am, he saw the first crude television images from the cabin. ‘‘I could not make out his facial features,’’ he wrote, ‘‘but I could tell from the way he moved that it was Yuri.’’

Leonov had been instructed not to initiate communications with Vostok unless given permission to do so and he duly remained silent. However, within moments, Gagarin radioed a request for details of his ‘‘flight path’’, keeping the orbital nature of the mission – for now – a secret from western ears. “The radio operator by my side did not realise his finger was depressing the button that opened the radio link… when he turned to speak to me,” Leonov wrote. The open mike broadcast the words of both Leonov and the radio operator straight up to Vostok. As soon as he heard Leonov tell the operator that “everything is going fine”, Gagarin responded with “Give my regards to Blondin”, fair-haired Leonov’s nickname.

The journey continued, cutting diagonally across the Pacific Ocean from volcanic Kamchatka’s land of fire and ice into darkness and toward the sleeping Americas. “The transition into Earth’s shadow,’’ Gagarin would explain to a Moscow press conference a few days later, “took place very rapidly. Darkness comes instantly and nothing can be seen. The exit from the Earth’s shadow is also rapid and sharp.’’ The unexpected suddenness of the shift from orbital daytime to nighttime led to some mutterings that the flight was a fraud; a testament, clearly, to how limited humanity’s understanding of space travel was at the time. Midway through the darkness, at 9:26 am, Vostok rose above the horizon of the Electronic Intelligence (ELINT) station on the Alaskan island of Shemya, giving the United States its first awareness that a Soviet man was indeed in orbit. Admittedly, American long-range radars had detected the R-7’s launch, but it was Shemya which confirmed without doubt that live dialogue was ongoing with an Earth-circling cosmonaut.

By 9:32 am, a stable orbit had been achieved – reaching a maximum apogee of 327 km and dipping to a perigee of 169 km, inclined 64.95 degrees to the equator – and, shortly thereafter, Gagarin began his passage across Hawaii, then out over the South Pacific. Following several requests, he was assured by the Khabarovsk station via long-range radio at 9:53 am that his orbit was satisfactory.

Minutes later, precisely on the hour, as he crossed the Strait of Magellan, Radio Moscow announced the electrifying news: “The world’s first spaceship, Vostok, with a man on board, was launched into orbit from the Soviet Union on 12 April 1961. The pilot-navigator of the satellite-spaceship Vostok is a citizen of the USSR, Flight Major Yuri Gagarin. The launching of the multi-stage space rocket was successful and, after attaining the first escape velocity and the separation of the last stage of the carrier rocket, the spaceship went into free flight on a round-the-Earth orbit. According to the preliminary data. . . ’’

The effect throughout the world, and particularly in the United States, was dramatic. Although it was known that the mission was underway virtually since its launch, and with certainty since Shemya confirmed Gagarin’s dialogue, and even though the imminence of the flight was not unexpected, the event sent shockwaves through the Kennedy administration. Before retiring to bed the previous night, the president had displayed a sense of foreboding and his science advisor, Jerome Wiesner, had gone so far as to prepare a statement for the press. Elsewhere, in Florida, NASA’s press officer John ‘Shorty’ Powers was awakened in the middle of the night by an alert journalist who had just heard of Gagarin’s launch. Powers, unfortunately, had not. With clear irritation in his voice and uttering words he would regret, he yelled down the phone: ‘‘What is this? We’re still asleep down here!’’ The journalist could not resist exploiting the figurative irony of Powers’ words. The United States and its own effort to put a man in space had indeed been caught off­guard. Next morning came the front-page headline: ‘Soviets Put Man In Space. Spokesman Says US Asleep.’

It was not only the world that expressed shock and disbelief at Gagarin’s achievement; even those closest to him – his family – had little or no inkling of what had happened. Only his wife Valentina knew that he would be flying into space, although he had told her that the mission was planned for 14 April, so as not to worry her. He informed his parents that he was going on a business trip and, according to his sister, would be travelling ‘‘very far’’. Upon learning of the flight, his mother’s reaction was to buy a train ticket to visit her daughter-in-law in Moscow and help care for their children. His father, meanwhile, was working on the collective farm near Klushino and, after hearing that someone called Major Yuri Alexeyevich Gagarin was in space, he responded that it could not be his son, who was ‘‘only a senior lieutenant”. The enormity of what his son had done became apparent when he headed to the local soviet for more information and was quickly pushed onto the phone with a Communist Party official in Gzhatsk. Within hours, Alexei Gagarin and his sons were answering an endless stream of calls from journalists within and beyond the Soviet Union.

High above Earth, at 10:25 am, just off the coast of western Africa, near Angola, Vostok’s retrorocket fired automatically to begin its re-entry into the atmosphere. Gagarin would never need to touch his controls, nor open the envelope, nor worry about the three magic numbers. Eight thousand kilometres still remained to be covered before landing back in the Soviet Union. However, all did not go according to plan. Ten seconds after retrofire, as intended, the four metal straps holding the instrument section and the capsule together severed, but an electrical cable failed to separate. The cable contained a thick bundle of wiring which provided power and data to the capsule.

For ten minutes, the two remained connected, with the unneeded instrument section trailing behind and causing the capsule to experience wild gyrations as high as 30 degrees per second. The incident, which was not revealed by the Soviets, but which Gagarin hinted at in his state-sanctioned report, was potentially disastrous. The capsule was weighted in such a way that it would naturally rotate to point the thickest part of its heat shield into the direction of travel. If the instrument section did not separate cleanly, the capsule could not assume its correct attitude and properly bear the brunt of re-entry heating. Gagarin could burn alive.

Fortunately, at 10:35 am, as the spacecraft’s meteoric descent path neared Egypt, the cable finally burned through and separated the instrument section. After slinging the capsule away with a spin so severe that Gagarin almost lost consciousness, it finally attained aerodynamic equilibrium with its heat shield positioned accurately. It was later revealed that the complications arose when a retrorocket valve failed to close properly, allowing some fuel to escape from the combustion chamber. As a result of this loss, the engine cut off a second too early, slowing Vostok at a less – than-expected rate and preventing a normal shutdown command from being issued. In the absence of this command, the engine’s propellant lines remained open and pressurised gas and oxidiser continued to escape through the nozzle and served to induce the wild gyrations.

Although the engine was ultimately cut off by a timer, its lack of delivered thrust caused the spacecraft’s control system to scrub the primary sequence to separate the capsule from the instrument section. The notes of the cosmonauts’ physician, Yevgeni Karpov, auctioned by Sotheby’s in March 1996, revealed his concern and, indeed, analysts have speculated that – had the world known of these problems – the Sixties space race might have slowed dramatically.

Gagarin himself made one brief reference to the incident. “The craft began to revolve and I told ground control about it,’’ he wrote. “The turning I had worried about soon stopped… ’’ Little else would be known or even suspected for almost three decades. However, in his testimony before the top-secret State Commission on 13 April, whose panel included both Korolev and Kamanin, he described feeling the intensity of the capsule’s oscillations and an audible sound of ‘cracking’, possibly from the structure of Vostok itself or from the expansion of the thermal cladding material. Through the Vzor, he saw the bright crimson glow of ionised gases rushing past the capsule and would later estimate that the G forces exceeded 10 and ‘‘data on the control gauges started to look blurry. . . starting to turn grey in my eyes’’.

Eventually, as he heard denser air whistling past the capsule and saw blue – not black – sky outside, he braced himself for ejection. By this point, Vostok had crossed back into Soviet territory, on the Black Sea coast, near Krasnodar. Gagarin’s ejection procedure was supposed to be fully automatic, triggered when the on-board sensors registered an outside atmospheric pressure consistent with an altitude of 7 km. He may, however, have felt that the oscillations were too severe to risk it and it would appear that he initiated the sequence manually. With a tremendous roar and rush of air, the hatch above his head blew away and the ejection seat’s rocket propelled him from the capsule at terrifying speed. Protected from frigid high – altitude temperatures of -30°C by his suit, Gagarin descended to Earth; his main parachute opened successfully, followed by its backup, and the First Cosmonaut hit the ground under two canopies.

The location at which his feet touched terra firma, in a field some 26 km south­west of the town of Engels in the Saratov region, near Smelovka, are today marked by a 12 m obelisk and plaque, inscribed with the legend ‘Y. A. Gagarin Landed Here’. The formal marker was placed there on 14 April 1961. However, the historic nature of the event had already led someone to erect a small commemorative signpost on the spot, instructing potential trespassers not to remove it and announcing the time of his landing as 10:55 am Moscow Time. Less than two hours had elapsed since Gagarin’s launch from Tyuratam.

Tractor driver Yakov Lysenko heard, but could not see, the ejection sequence as a loud ‘crack’ in the sky. Seconds later, he saw the Vostok capsule descending under its own automatically-deployed parachute and immediately returned to Smelovka to raise the alarm. A hastily-assembled search party was greeted by what Lysenko would later describe as a ‘‘very lively and happy’’ Gagarin, who identified himself to them as ‘‘the first space man in the world’’. Farmer Anna Takhtarova also recalled the strange sight of the orange-clad cosmonaut telling them not to be afraid and asking for a telephone to call Moscow.

Shortly thereafter, the Soviet military, under General Andrei Stuchenko, arrived in force to take over the recovery effort. Gagarin had been promoted to the rank of major during his flight and was greeted as such by one of the local officers, Major Gasiev. “It was a complete invasion force,” Yakov Lysenko said later of the military’s arrival. “They didn’t allow us to get too close. They’re very strange people.” It was understandable, at least from Stuchenko’s perspective. He had already been told in no uncertain terms by an anonymous Kremlin official the previous day that on his head, literally, lay the responsibility to safely recover Gagarin. Stuchenko obviously wanted to leave nothing to chance.

The Vostok capsule hit the ground a couple of kilometres from the cosmonaut himself, since his high-altitude ejection had caused them to drift apart. For years, the spacecraft’s landing site was officially one and the same with Gagarin’s own, to avoid FAI suspicions that both had not touched down together. However, the Vostok site is known with certainty, thanks to a group of children who happened to be playing in a meadow near the banks of a tributary to the Volga River. They saw the capsule land. Schoolgirls Tamara Kuchalayeva and Tatiana Makaricheva described the dent it left in the soft earth and related how the boys clambered inside and began handing out and trying the tubes of space food. “Some of us were lucky and got chocolate,’’ Makaricheva recalled of the unusual mid-morning snacks. “The others got mashed potatoes. I remember tasting some [of this] and spitting it out.’’ Kuchalayeva agreed that she would not eat it again. Gagarin’s 108-minute adventure, it seemed, made him a hero not only for being the first to survive space, but also for being the first to survive the tastelessness of space food.

Monkeys to Men

HEROES

An expectant hush descended over the journalists in the conference room of the Dolley Madison House, opposite Lafayette Park in downtown Washington, DC. It was the afternoon of 9 April 1959. Backstage, clad in civilian suits, two of them wearing bow ties, stood the United States’ first team of astronauts – the ‘Mercury Seven’. At 2:00 pm Eastern Standard Time, the presiding officer, Walt Bonney, finally spoke. ‘‘Ladies and gentlemen,’’ he began, ‘‘in about 60 seconds, we will give you the announcement you have all been waiting for: the names of the seven volunteers who will become the Mercury astronaut team.’’

Those volunteers – Scott Carpenter, Gordo Cooper, John Glenn, Virgil ‘Gus’ Grissom, Wally Schirra, Al Shepard and Donald ‘Deke’ Slayton – had been chosen after three months of careful screening of 110 experienced combat and test pilots. The list had gradually been reduced through a series of invasive and, in many cases, degrading medical and psychological evaluations. The pilots had been split into three groups, two of which were summoned secretly to the Pentagon in January 1959 for briefings on the effort to send a man into space. Afterwards, the 69 candidates from the first two groups had been asked if they wished to volunteer for further tests and, to the great surprise of many on the selection board, more than 90 per cent of them agreed. The third group, unneeded, was never called up. It had been assumed that the pilots would be so entrenched in their military careers that shifting to a civilian space project would hold little appeal. Already, some of the United States’ most accomplished test pilots – Chuck Yeager among them – had poured scorn on the idea of rocketing men into space atop converted ballistic missiles. Even Al Shepard’s father (himself a former military officer) expressed concern that his son had made the wrong decision. He would eat his words years later when Shepard not only became the first American in space, but also the fifth man to walk on the Moon.

Sixty-nine candidates rapidly dwindled to just thirty-two, who reported to Randy Lovelace’s aerospace medicine clinic in Albuquerque, New Mexico, for additional evaluations. Over the course of a week, every spot on the bodies of Shepard and the

Monkeys to Men

The Mercury Seven. From the left are Gordo Cooper, Wally Schirra (partially obscured), Al Shepard, Gus Grissom, John Glenn, Deke Slayton and Scott Carpenter. Behind them is a Mercury capsule, which all of them but Slayton flew into space.

others was sampled, measured, poked and prodded, with scarcely a muscle, bone or gland left untouched. Throats were scraped, stool and semen samples taken, electricity zapped into hands and intensely uncomfortable ‘steel eels’ inserted into rectums. Wally Schirra would later call Lovelace’s tests ‘‘an embarrassment, a degrading experience… sick doctors working on well patients’’.

Survivors of the clinic devised a tradition of inviting the newer candidates to dinner at a local Mexican restaurant. At one of these gatherings, the veterans each had at their feet a jug of urine, which they had been obliged to collect for medical purposes during their stay. One evening, accidentally, Gus Grissom knocked over his jug, but, thanks to the quick-thinking crowd of test pilots, was provided with a ready solution: to order more beer. Several rounds, and a number of trips to the lavatory, later, Grissom’s jug had its required amount of urine…

Still more tests followed at the Aeromedical Laboratory of Wright-Patterson Air Force Base in Dayton, Ohio, where the pilots withstood cold water pumped into their ears, sat for hours in overheated saunas, endured soundproofed and darkened isolation rooms, blew up balloons until they were out of breath, walked on treadmills until their heart rates soared to 180 beats per minute and were photographed from every conceivable angle and into every conceivable orifice. Many perceived the whole thing as excessive and a waste of time. “I’d flown combat missions and done operational test flying for 17 years by that point,’’ wrote Deke Slayton. “The fact that I’d survived should have told them all they needed to know about stress. At least putting me in the blackout chamber, they let me catch a nap!’’

Not only were the selectors looking for the most physically unbreakable men, they were also scrutinising their reactions to the tests and the testers. Would they crack, psychologically, under the unknown stresses imposed by the mysterious space environment? Personality questions prompted them to explore their individual motivations for wanting to become astronauts, their concerns about their health, their frustrations, their ‘thoughts’ and even whether their desires to fly jets and rockets arose from feelings of sexual inadequacy. The seven men eventually chosen, in addition to their combat and test flying credentials, were all highly intelligent – with IQs of between 131 and 141 – but, said psychologist George Ruff, all were ‘‘oriented toward action, rather than thought’’.

EUPHORIA, DISAPPOINTMENT… AND A BOSS

The weeks after Friendship 7 would be filled with euphoria over lohn Glenn’s achievement, tempered with disappointment over the fate of Deke Slayton, the man meant to follow him into orbit on the next Mercury-Atlas mission.

Relief at Glenn’s safe return was evident on many faces, including that of lawyer Leo D’Orsey, who had personally endorsed a $100,000 cheque for Annie to cover her husband’s life insurance in the event of a disaster. (Despite D’Orsey’s attempts to purchase million-dollar policies, the astronauts were uninsurable.) Now, thankfully, it did not need to be cashed. “Boy, am I glad to see you!’’ D’Orsey told Glenn later. Elsewhere, Henri Landworth, former manager of Cocoa Beach’s Starlight Motel and later of the Holiday Inn frequented by the astronauts, had made a 400 kg cake, shaped like Friendship 7. He even rigged up an air-conditioned truck to prevent it from spoiling. After Glenn sampled a slice, Landwirth told him that it had been made in time for the original lanuary launch date and was a month old!

At Cape Canaveral, President Kennedy awarded the astronaut and Bob Gilruth with NASA’s Distinguished Service Medal, then flew America’s newest hero to Washington, DC, aboard Air Force One to address Congress and, later, enjoy a tickertape parade through the streets of New York. Whilst airborne, Kennedy introduced his four-year-old daughter Caroline to Glenn. Clearly accustomed to the flights of Ham and Enos, the little girl’s disappointed reaction was: “But where’s the monkey?’’

Like Shepard, Glenn was impressed by the president’s enthusiasm and passion for the space programme. “He believed… that it was not just a scientific journey,’’ the

Подпись: 130 Monkeys to Men

EUPHORIA, DISAPPOINTMENT... AND A BOSS

John Glenn experiences his first taste of public adoration, with President Kennedy at his side.

 

astronaut wrote in his memoir, “but a source of inspiration that could motivate Americans to pursue great achievements in all fields.” Glenn had been accorded the rare privilege of addressing a joint session of Congress, which he did on 26 February, before the tickertape parade through New York. Four million people reportedly lined the streets to greet him and his speech before Congress led some columnists, including Arthur Knock and James Reston of the New York Times, to remark on his suitability for politics and his embodiment of “the noblest human qualities”. One group of Republicans from Nevada even called upon him to run for the presidency. Glenn would admit at the time that his interests were apolitical, but a cultivated friendship with, among others, Attorney-General Bobby Kennedy would gradually direct him towards a career in the Senate.

His diplomatic abilities were also put to the test in May 1962, when he was invited, along with Vostok 2 cosmonaut Gherman Titov, to address the Committee on Space Research, part of the third International Space Symposium, in Washington, DC. Glenn described Titov as “cordial but forceful and thoroughly indoctrinated. . . charts and photographs had supplemented my presentation, but not his; he followed the Soviet line that disarmament would have to precede full sharing of information”. Nonetheless, he hosted the cosmonaut during his visit to Washington, even debating the existence of God in front of journalists.

Titov’s standpoint was that he did not see anyone in space – offering clear “proof for the communist position’’ that such a deity did not exist – but Glenn responded that ‘his’ God was not so small that he expected to “run into Him a little bit above the atmosphere”. Diplomacy was extended still further when the Glenns invited the Titovs to a barbecue at their house; after initially refusing, the cosmonaut’s delegation changed its mind at the last moment. The Glenns, unprepared, were forced to ask their neighbours for spare steaks and send their police escorts to buy vegetables. Meanwhile, Al Shepard picked up the Titov delegation and bought the Glenns some time by taking a few wrong turns on his way to Arlington. . .

Not only was Glenn deemed valuable to the nation, but so too was Friendship 7, which, despite some discolouration, was in remarkable condition and today resides in the National Air and Space Museum in Washington, DC. In 1962, however, it undertook what was popularly nicknamed its ‘fourth orbit’: a global tour of 17 countries before eventually being installed in the Smithsonian, close to the Wright Brothers’ original aircraft and Charles Lindbergh’s ‘Spirit of St Louis’.

As Glenn basked in his new-found fame, 38-year-old Donald Kent ‘Deke’ Slayton fought a losing battle to fly his own Mercury-Atlas mission, tentatively scheduled for April 1962. He had already picked a name for his capsule – ‘Delta 7’, the fourth letter in the Greek alphabet for the fourth manned Mercury mission – which he would describe in his autobiography as ‘‘a nice engineering term that described the change in velocity’’. Slayton’s own velocity, both in spacecraft and high-performance aircraft, would decline markedly, thanks to a minor, yet persistent, heart condition, known as ‘idiopathic atrial fibrillation’. This took the form of occasional irregularities of a muscle at the top of the heart, caused by unknown factors and extremely rare in highly-fit thirtysomethings like Slayton. It had first arisen during a centrifuge run at Johnsville in August 1959, when physicians noticed traces of sinus arrhythmia, which Bill Douglas later wrote “wasn’t uncommon in healthy young men and… the kind of thing that often went away with exertion’’. After the run, however, it was still present, prompting Douglas and his team to undertake a clinical electrocardiogram at the Philadelphia Navy Hospital.

They concluded that Slayton had a ‘flutter’ in his heartbeat, although, in 1959, the astronaut himself ‘‘had no idea how much of a problem it was’’. Further tests at the Air Force’s School of Aviation Medicine in San Antonio, Texas, verified that the condition was of no consequence and should not influence Slayton’s eligibility for a spaceflight. Douglas informed Bob Gilruth, who briefed NASA Headquarters on the issue late in 1959, as well as the Air Force Surgeon-General, who advised that no further action was necessary. The ‘Slayton File’, for a time, lay dormant.

The problem resurfaced a little over two years later, a week before Friendship 7’s launch, when speculation arose that John Glenn had a heart murmur. Apparently, wrote Slayton, the call to Bill Douglas came from Air Force physician George Knauf, attached to NASA Headquarters, and had originated from ‘‘a source higher than the Department of Defense’’. Douglas denied that Glenn had a problem, but effectively opened a can of worms. Knauf asked next if Glenn’s backup, Scott Carpenter, had a heart murmur: again, the response was negative. Then Douglas, to reinforce the point that the matter was of little consequence, revealed that Slayton had long been known to have a minor condition. He expected this to be the end of the matter. It wasn’t.

Back in 1959, flight surgeon Larry Lamb had examined Slayton at Brooks Air Force Base in San Antonio and had become convinced the heart fibrillation should disqualify him from the selection process. ‘‘He hadn’t said so in 1959,’’ wrote Slayton, ‘‘but he said so now. I don’t think it was anything personal – this was just his medical opinion.’’ Lamb’s judgement was very much a voice in the wilderness, but unfortunately he also happened to be Lyndon Johnson’s cardiologist, and in the spring of 1962 began to question the astronaut’s suitability for flight. Three weeks after Friendship 7, Jim Webb reopened Slayton’s medical files and the astronaut and Douglas were summoned to the office of the surgeon-general of the Air Force in Washington, DC. A panel of military physicians signed him off as fit to fly, a decision endorsed by the Air Force’s chief of staff, General Curtis LeMay.

For Webb, though, it was not enough. The Secretary of the Air Force, Eugene Zuckert, suggested that a civilian panel of physicians should also examine Slayton at NASA Headquarters. On 15 March, less than two months before launch, Slayton was poked and prodded and had his heart monitored by Proctor Harvey of Georgetown University, Thomas Mattingley of the Washington Hospital Center and Eugene Braunwell of the National Institutes of Health. In a turnaround which, in Slayton’s own words, would leave him ‘‘devastated’’, Deputy Administrator Hugh Dryden entered the room and told him, point-blank, that he was off the flight. None of the physicians had found a specific medical reason to keep him off Delta 7, but their consensus was that if NASA had pilots ‘without’ his condition, one of them should fly the mission instead.

Years later, Slayton would refute theories that Lyndon Johnson’s annoyance over the Annie Glenn incident had anything to do with the decision, but certainly felt it was a political move. “NASA knew it would have to publicly disclose my heart condition prior to my flight,” he wrote. “There would be medical monitors at tracking stations all over the world who wouldn’t know how to react otherwise. Everybody expected this to be a big deal. NASA would be opening itself up to a lot of medical second-guessing.” Bill Douglas felt that problems could arise if Slayton started fibrillating on the pad – “do you scrub the launch or go ahead?’’ – but he, Bob Gilruth and Walt Williams had confidence that he was the best person to follow John Glenn. All three men were prepared, personally, to take the heat, but Jim Webb’s fear that it could trigger adverse headlines for the agency drew a line in the sand. “It didn’t matter that a whole lot of doctors thought I didn’t have a problem,’’ Slayton wrote of Webb’s actions. “He was only going to listen to the few who did.’’

In Webb’s mind, an Atlas abort could subject the astronaut to acceleration loads as high as 21 G and conjured the very real possibility that Slayton, dehydrated and perhaps fibrillating, could die as a result. The impact on NASA, on President Kennedy’s promise to land a man on the Moon before the end of the decade and on the ongoing contest with the Soviets could be profound.

The next day, 16 March, the now-grounded, and furious, astronaut was forced to sit through a lengthy press conference, in which the minutiae of the case were examined. Hugh Dryden remarked that, despite the decision, Slayton might remain eligible for future flights. One journalist asked if the problem had been caused by stress, to which Slayton responded no, and further that he did not even know about it until he had been hooked up to an electrocardiogram in 1959. The most stressful part of the space business, he explained caustically and with more than a hint of sarcasm, was “the press conference after the flight’’. Bill Douglas’ own departure from NASA within days of the announcement, to return to the Air Force, was leapt upon by some journalists as ‘evidence’ of his bitterness over Slayton’s treatment. In truth, Douglas was already at the end of a three-year detachment to NASA and his transfer had been in the works since mid-1961.

Despite his grounding, Slayton did not give up on flying. ‘‘I made some changes in my lifestyle,’’ he wrote, ‘‘gave up drinking, started working out more regularly – quit doing everything that was fun, I guess!’’ Thanks to Bill Douglas, he also secured an examination by Dwight Eisenhower’s cardiologist, Paul Dudley White, in June 1962. White advised him that two-thirds of people with his condition would die young, whilst the remainder would probably never know they had it and might never be affected. The verdict: ‘‘Young man, you’re going to live a long time.’’ (Slayton lived to be 69.) However, White’s report, which highlighted that Slayton did not appear to have a problem, also advised that if astronauts were present without the condition, it would be preferable to assign them in his stead. As it became clear that he would not draw any of the remaining Mercury-Atlas flights, Slayton turned his gaze to the two – man Gemini project, only to be told by Bob Gilruth that his condition would make him a ‘‘hard sell’’ to senior management. Shortly thereafter, the Air Force decreed that he no longer met the qualifications for a Class I pilot’s licence – effectively, he could no longer fly solo – and, at the end of November 1963, Slayton tendered his resignation from the service.

Although he would eventually get his ride into space, it would not come until July 1975, when he was 51 years old. A lesser man might have thrown in the towel and departed NASA for pastures new, but not Slayton. With no guarantee that he would ever fly, he decided to stay and in the summer of 1962, as the agency prepared to expand its astronaut corps by picking nine new pilots, he was appointed as Co­ordinator of Astronaut Activities. Initial plans to bring in a manager from the outside to oversee the corps were quashed by the astronauts themselves. “What we wanted the least,” wrote Wally Schirra, “was somebody who would outrank us and issue orders in a military way. We wanted someone who knew us, who trained with us. Deke was the one and only choice.” During the next few years, as America pushed for the Moon, Slayton, though non-flying, would be all-powerful within the astronaut corps, deciding the career paths of the men who would someday walk on the lunar surface. . . and those who would not.