Category Escaping the Bonds of Earth

Disaster, Recovery, Triumph

“FIRE”

In a barren, disused area of Cape Canaveral stands a gaunt, concrete-and-steel hulk which once formed the launch platform of Pad 34. Four decades ago, it served as the starting point for the first Apollo mission, an 11-day trek into low-Earth orbit to demonstrate the capabilities of the spacecraft that would one day deliver men to the Moon. Today, overgrown with bushes, weeds and a few wild pepper trees, it slowly decays in the salty air. A faded ‘Abandon in Place’ sign adorns one of its legs. Near its base are a pair of plaques, memorialising a far darker and more tragic event. The first reads simply: ‘Launch Complex 34 Friday 27 January 1967 1831 Hours’ and dedicates itself to the first three Apollo astronauts, Virgil ‘Gus’ Grissom, Ed White and Roger Chaffee. The second pays tribute to their ‘ultimate sacrifice’ that January evening, long ago. Nearby are three granite benches, one in honour of each man.

Every year, without fail, their families are invited by NASA to visit the spot and reflect upon the disaster which befell them that Friday. It is a time of year which has become synonymous with tragedy in America’s space programme; the losses of Challenger and Columbia having occurred within the very same week in 1986 and 2003. Indeed, 27 January 1967 marked a line in the sand, as the seeming good fortune and achievement of Project Gemini gave way to the stark and brutal realisation that reaching the Moon before the end of the decade was by no means assured and, technically and literally, our closest celestial neighbour remained a long way off. It would be a bad year for both the United States and the Soviet Union and by its close no fewer than six spacefarers would be dead: three killed in Apollo 1, a cosmonaut during his return to Earth and two astronauts in aircraft and car accidents.

Gene Cernan, for whom 1966 ended brightly with assignment to the Apollo 2 backup crew, recounted that 1967 was such a rotten year that his wife, Barbara, did not even feel able to write her usual Christmas letters to friends. It should have been quite different. NASA’s plans called for as many as three Apollo missions, the first (led by Grissom) employing a spacecraft design known as ‘Block 1’, capable only of

reaching Earth orbit. The others, commanded, respectively, by Jim McDivitt and Frank Borman, would utilise a more advanced Block 2 type, which possessed the navigational, rendezvous and docking equipment needed for lunar expeditions. “There were hundreds of differences between the two,” wrote Deke Slayton, “the major one being that Block 1 vehicles didn’t have the docking tunnel that would allow you to dock with a lunar module.’’

Each mission, though, would find common ground in that it would evaluate what Grissom had already described as a machine infinitely more complex than Mercury or Gemini; a machine which came in two parts. The ‘command module’, firstly, was a conical structure, 3.2 m high and 3.9 m across its ablative base, and would provide its three-man crews with 5.95 m3 of living and work space whilst aloft. Brimming with reaction controls, parachutes, propellant and water tanks, cabling, instrumen­tation and controls, it would truly live up to its name as the command centre for future voyages to the Moon. Both it and the second section, the ‘service module’, were built by North American Aviation of Downey, California – today part of Boeing – and the latter consisted of an unpressurised cylinder, 7.5 m long and 3.9 wide, housing propellant tanks, fuel cells, four ‘quads’ of manoeuvring thrusters, an S-band communications antenna, oxygen and water stores and the giant Service Propulsion System (SPS) engine. The latter, some 3.8 m long and fed by a propellant mixture of hydrazine and unsymmetrical dimethyl hydrazine with an oxidiser of nitrogen tetroxide, would provide the impulse for inserting Apollo into and removing it from lunar orbit. It was a critical component without a backup and this demanded that its design be as simple as possible: its propellants, pushed into the combustion chamber by helium, were ‘hypergolic’, meaning that they would burn on contact, with no need for fuel pumps or an ignition system. The propellants would flow so long as the valves were held open and the valves were designed to be extremely reliable. The command and service modules would remain connected throughout a mission, with the latter jettisoned just minutes before re-entry.

Evaluating this machine for the first time was Grissom’s responsibility. He and his crew were tasked with an ‘open-ended’ mission, lasting anywhere from six orbits to 14 days. In his book on lost and forgotten Apollo missions, spaceflight historian Dave Shayler wrote that it was certainly the crew’s desire to fly for as long as practical – to extract as much valuable engineering data from the spacecraft as they could – but admitted that the Block 1 vehicle was almost certainly incapable of supporting operations for longer than 14 days. It will never be known what duration Apollo 1 might have achieved, but it seems doubtful that it could have seriously threatened the 14-day record established by Frank Borman and Jim Lovell a year earlier. Indeed, in the weeks before 27 January, Gus Grissom joked darkly that as long as the crew returned from orbit alive, he would consider the mission a success. His fatalistic outlook and lack of confidence in Block 1 was shared by many of his fellows within the astronaut corps.

Their launch, according to Flight Director Chris Kraft at the Apollo News Media Symposium, held in mid-December 1966, was targeted for late February or early March of the following year. As the event drew nearer, this date was refined. By the end of January, it was set for 21 February. Internally, the mission was known as

‘Fire’

‘Apollo-Saturn 204’ (AS-204) and Grissom, White and Chaffee would have become the first humans to ride Wernher von Braun’s mighty Saturn IB booster, a two-stage behemoth which, in just 11 minutes and 30 seconds, would have injected their 20,400 kg spacecraft into an orbit of 136 x 210 km. (The ‘204’ indicated that the launch vehicle was the fourth production unit of the Saturn 1B type. Despite this nomenclature, the crew had successfully pushed to rename their flight ‘Apollo 1’.)

Grissom and his backup, Wally Schirra, light-heartedly dubbed the Saturn 1B ‘‘a big maumoo’’. Its first stage, known as ‘S-IB’, was 25.5 m tall and 6.6 m wide and would have boosted Apollo 1 to an altitude of 68 km under the combined thrust of its eight H-1 engines. Next, the ‘S-IVB’ second stage, fitted with a single J-2 engine, would have completed the climb into orbit. The latter, built by the Douglas Aircraft Company, would also form the third stage of the Moon-bound Saturn V booster. Minutes after achieving Earth orbit, Ed White, the senior pilot of Apollo 1, would have unstrapped and headed into the command module’s lower equipment bay to begin setting up cameras and scientific hardware. Two and a half hours into the flight, he and pilot Roger Chaffee would have returned to their seats as Grissom separated Apollo 1 from the S-IVB.

Under Grissom’s control, the spacecraft would have maintained tight formation with the spent stage as White and Chaffee employed a battery of cameras to record the venting of its residual liquid oxygen and hydrogen propellants. Much of this work would be crucial for subsequent, Moon-bound missions, which would involve a third Apollo component, the lunar module, housed inside the upper part of the S – IVB and extracted by the command and service module shortly after launch. No rendezvous manoeuvre with the stage was planned on Apollo 1, however, but several tests of the SPS engine were timetabled. In total, it would have been fired on eight occasions: three times by Grissom and White and twice by Chaffee, with the astronaut responsible for ‘flying’ each burn stationed in the commander’s seat. The other crew members would respectively fulfil ‘navigation’ and ‘engineering’ roles. The first pair of SPS firings would have occurred during Apollo 1’s second day in orbit, with the remainder being executed at roughly 50-hour intervals throughout the rest of the mission.

Additionally, the crew was assigned one of the largest complements of scientific, photographic and medical experiments ever carried into orbit. During their long flight, they would have operated instruments to monitor aerosol concentrations in the command module’s cabin, carried out synoptic terrain photography of diverse targets ranging from the coast of Africa to the Mississippi River Mouth and Oyster Bay in Jamaica to the South China Sea and observed a range of meteorological and marine phenomena, from cloud eddies to dust storms and smog-laden cities to ocean currents. Other photographic tasks and medical investigations – an in-flight exerciser, a photocardiogram to monitor heart function and an otolith ‘helmet’ to track the effects of weightlessness on the balance mechanism of the inner ear – would have filled much of their time. Moreover, they would have been required to evaluate every aspect of Apollo itself, from the performance of its rudimentary ‘toilet’ to its general habitability.

Had their mission gone ahead and run to its maximum length, the three men would have returned to Earth on 7 March, separating from their service module a few minutes before plunging, base-first, into the upper fringes of the atmosphere. The command module would then have borne the brunt of re-entry heating and finally, beneath a canopy of three red-and-white parachutes, would have splashed into the Pacific some 330 hours after launch. Assuming nothing untoward happened, the flight of Apollo 1 would clear the way for the first mission of the Block 2 spacecraft and, eventually, for a lunar landing, possibly as soon as late 1968.

Thanks to Wally Schirra, there would only be one manned Block 1 flight. Originally, NASA wanted to virtually duplicate Grissom’s mission with Apollo 2, featuring Schirra and rookie astronauts Donn Eisele and Walt Cunningham. However, the man who had commanded the world’s first space rendezvous in December 1965 wanted nothing to do with it. “I argued it made no sense to do a repeat performance,” he wrote, “and I succeeded in getting the mission scrubbed.’’ (In fact, Walt Cunningham, in a 1999 oral history for NASA, speculated that Deke Slayton himself hoped to command Apollo 2, but was overruled. Slayton made no reference to this in his autobiography.) Regardless of Schirra’s involvement, the hands of fate had already turned on Apollo 2 by October 1966, when the propellant tanks inside one of the service modules exploded during a ground test. Although it had not been assigned to Apollo 2, it made sense to keep downstream flights on schedule, replace it with Schirra’s service module and cancel the ‘repeat performance’. Whether he liked it or not, Schirra was told on 15 November that Apollo 2 was no more.

In winning his battle to get rid of the second Block 1 mission, Schirra had shot himself in the foot because he, Eisele and Cunningham were reassigned in early December as Grissom’s new backup team. Schirra was reluctant to accept the change – he wanted a mission of his own that would provide a challenge and certainly did not want to serve on another backup crew – but Slayton and Grissom eventually talked him around. (He would complain to Tom Stafford, however, that Slayton had ‘‘screwed him’’.) Meanwhile, astronauts Jim McDivitt, Dave Scott and Rusty Schweickart, who had served as Apollo 1’s backups since March 1966, would now become the prime crew of a ‘new’ Apollo 2, which would test the Grumman-built lunar module in Earth orbit on a mission involving two Saturn 1Bs. McDivitt, to be fair, had been working on the lunar module for more than a year and Scott had rendezvous experience, so it made sense to assign them, rather than Schirra’s team, to this mission. Next up would be Frank Borman, Mike Collins and Bill Anders on Apollo 3, tasked with the first manned flight of the Saturn V to a high Earth orbit, reaching an record-breaking apogee of 6,400 km. Had the 27 January tragedy not occurred, these missions might have followed Apollo 1 in the summer and autumn of 1967, setting the stage for a lunar landing the year after.

Schirra’s wish to cancel the second Block 1 flight was not just based on a whimsical reluctance to fly what he perceived to be a pointless mission. It also had much to do with his lack of confidence in the Block 1 design, which many astronauts considered sloppy and unsafe. ‘‘The craft was like an old friend,’’ Time magazine told its readers on 3 February 1967. It was nothing of the sort. Since his assignment

The Apollo 1 command and service module during testing, with the Service Propulsion System (SPS) engine and thruster quads clearly visible. Apollo 1 would have been the first and only flight of the ‘Block 1’ design – a design disliked and distrusted by both Gus Grissom and Wally Schirra.

to Apollo 1, Gus Grissom had spent months overseeing poor performance and low standards on the part of North American in their efforts to get the Block 1 hardware ready for space. Unlike Gemini, where they could approach James McDonnell himself if issues arose, the North American set-up was far larger, more impersonal and little heed was paid to the astronauts’ concerns. It “was a slick, big-time bunch of Washington operators,’’ wrote Tom Stafford, “compared to the mom-and-pop operation at McDonnell.’’ Some astronauts felt the technicians were more worried about their free time than with building a Moonship. Even NASA’s Apollo manager Joe Shea remarked that, after receiving the two-billion-dollar contract, North American had thrown a party and made hats with ‘NASA’ printed on them, albeit with the ‘S’ replaced by a dollar sign…

In fact, as 1966 wore on, the number of problems with ‘Spacecraft 012’ – the command and service module assigned to Grissom’s mission – was so high that technicians were having trouble tending to them all. Riley McCafferty, responsible for updating the simulators, noted that at one point more than a hundred modifications awaited implementation. Plans to ship Spacecraft 012 to Cape Kennedy in August 1966 were postponed by three weeks when problems arose with a water glycol pump in the command module’s environmental control system. Upon arrival in Florida, more than half of the engineering work assigned to North American remained incomplete, together with other serious deficiencies: a leaking SPS engine, coolant problems, computer software that ‘‘never quite worked right’’ and faulty wiring.

It did not inspire confidence in Grissom’s crew, who presented Shea with a half – joking photograph of themselves bowed in prayer over their spacecraft. ‘‘It’s not that we don’t trust you,’’ they told him, ‘‘but this time we’ve decided to go over your head!’’ Repairs and rework eliminated any hope of launching Apollo 1 before the end of the year and a February 1967 target became unavoidable. On 22 January, just before leaving his Houston home for the last time, Grissom was so angry with the problems that he plucked a grapefruit-sized lemon from a tree in his backyard, flew it to the Cape in his baggage and hung it over the hatch of the Apollo simulator.

Things had seemed quite different the previous March, when Grissom, White and Chaffee, ‘‘the coolest heads in the business,’’ according to Bob Gilruth, had been named as the first Apollo crew. Together with their backups, McDivitt, Scott and Schweickart, they had closely monitored the spacecraft’s progress throughout 1966, spending virtually every waking hour in Downey, renting rooms in a nearby motel and even sleeping in bunks near the production line from time to time. ‘‘Gus and Jim were the first two guys available to move from Gemini to Apollo,’’ wrote Deke Slayton, ‘‘which is why I assigned them. Gus would keep an eye on the command and service module, while Jim would start following development of the lunar module.’’

To be fair, North American had faced immense technical challenges of its own. One of these was NASA’s mandate that the command module should operate a pure oxygen atmosphere – a dangerous fire hazard, admittedly, but infinitely less complex than trying to implement an oxygen-nitrogen mixture which, if misjudged, could suffocate the crew before they even knew about it. In space, the cabin’s atmosphere would be kept at a pressure of about a fifth of an atmosphere, but for ground tests would be pressurised to slightly above one atmosphere. This would eliminate the risk of the spacecraft imploding, but at such high pressures there remained the danger that anything which caught fire would burn almost explosively. At an early stage, North American had objected to the use of pure oxygen aboard Apollo, but NASA, which had employed it without incident on Mercury and Gemini, overruled them.

The selection of pure oxygen was not made lightly. NASA engineers had long been aware that a two-gas system, providing an approximately Earth-like atmosphere of some 20 per cent oxygen and 80 per cent nitrogen, pressurised to one bar, would reduce the risk of fires. Further, having a gaseous mixture of this type would avoid many of the physiological effects of pure oxygen, such as eye irritation, hearing loss and a clogging of the chest. However, at the time, the complexities of building a system which could mix and monitor these gases would have added to the spacecraft’s weight, making it intolerably heavy. Other complications included the astronauts’ space suits, which were pressurised to 0.31 bars. “To walk on the Moon,’’ wrote Deke Slayton, “you needed to get out of the spacecraft… and with a mixed-gas system you’d have to pre-breathe for hours, lowering the pressure and getting the nitrogen out of your system so you didn’t get the bends. Of course, if there was a real emergency and you had to use the suit, you’d really have been in trouble.’’

Other worries surrounded the craft’s hatch, a complex affair which actually came in two cumbersome segments: an inner piece which opened into the command module’s cabin, overlaid by an outer piece. North American had proposed a single­piece hatch, fitted with explosive bolts, which could be swung open easily in an emergency; NASA, however, argued that this might increase the risk of it misfiring on the way to the Moon. By adopting an inward-opening hatch, cabin pressure would keep it tightly sealed in flight. . . but notoriously tough to open on the ground. These two factors – a pure oxygen atmosphere and an immovable hatch – coupled with a mysterious ignition source would spell death for the Apollo 1 crew.

That Grissom was beginning to feel a strong sense of foreboding about the mission is reflected in his propitious comments in the weeks preceding the 27 January test – most famously, his declaration that “if we die, we want people to accept it… this is a risky business’’. He had already told colleagues that, if an accident did occur in the programme, it would probably involve him. Also, for the first time, he began to take the frustrations of work home with him. “When he was home,’’ recalled his wife, Betty, “he normally did not want to be with the space programme. He would rather be just messing around with the kids, but he was uptight about it.’’ Others, including Walt Cunningham, added that Spacecraft 012 “just wasn’t as good as it should have been for the job of flying the first manned Apollo mission’’.

It was with an air, perhaps, of scepticism and contempt that Grissom, White and Chaffee, clad in their white space suits, crossed the gantry at Pad 34 and eased themselves into their seats early on the fateful afternoon of 27 January. All three were described by NASA secretary Lola Morrow – who had herself nicknamed Apollo as ‘Project Appalling’ – as unusually subdued and in no mood for the so – called ‘plugs-out’ test. By this point, the backup crew had changed to Wally Schirra,

Donn Eisele and Walt Cunningham, who had themselves spent the previous night aboard the spacecraft for a ‘plugs-in’ test, with the vehicle still entirely dependent upon electrical power from the ground and the hatch open. Schirra liked Spacecraft 012 no more than did Grissom. After emerging from his own test, he took his friend to one side and told him to ‘‘get outta there’’ if he sensed even the slightest glitch. ‘‘I didn’t like it.’’

Communications between the spacecraft and the nearby blockhouse, manned by rookie astronaut Stu Roosa, caused problems almost from the outset. Grissom was so frustrated that he had even asked Joe Shea, at breakfast that morning, to climb into the cabin with them and witness, first-hand, from a manager’s perspective, how bad the problems were. Shea weighed up the pros and cons of rigging up a headset and somehow squeezing himself in shirtsleeves into Apollo 1’s lower equipment bay, beneath the astronauts’ footrests, but ultimately decided against it. Even Deke Slayton, who was based in the blockhouse that day, considered sitting in with them, but decided to remain where he was and help monitor the test.

Grissom was first aboard, taking the commander’s seat on the left side of the cabin, and almost as soon as he had hooked his life-support umbilicals into the oxygen supply, he noticed a peculiar odour. It smelled, he said, like sour buttermilk, and technicians were promptly scrambled to the spacecraft to take air samples. Nothing untoward was found and Roger Chaffee entered next, taking his place with the communications controls on the right-hand side. Finally, Ed White slid into the centre couch. With all three men aboard, the command module’s inner and outer hatches were closed and sealed and finally the boost cover to protect it from the exhaust of the Saturn 1B’s escape tower was secured. Next, pure oxygen was steadily pumped into the cabin.

Throughout the afternoon, a multitude of niggling problems disrupted and delayed things. A high oxygen flow indicator periodically triggered the master alarm and spotty communications between Apollo 1 and Roosa grew so bad that Grissom barked in exasperation: ‘‘How are we gonna get to the Moon if we can’t talk between two or three buildings?’’ One such problem, which arose at 4:25 pm, turned out to be caused by a live microphone that could not be turned off. The NASA test conductor, Clarence Chauvin, recalled that communications were so poor that his team could barely hear the astronauts’ voices. Eventually, the simulated countdown was put on hold at 5:40 pm. Forty minutes later, after more communications headaches, controllers prepared to transfer the spacecraft to internal fuel cell power. . . and the countdown was held again.

Suddenly, with no warning, controllers noticed the crew’s biomedical readings jump, indicating increased oxygen flow in their space suits. At the same time, around 6:30:54 pm, other sensors recorded a brief power surge aboard Apollo 1. Ten seconds later came the first cry from the spacecraft. It was Roger Chaffee’s voice: ‘‘Fire!’’

Seated in the blockhouse, Deke Slayton glanced over at a monitor which showed Apollo 1’s hatch window, normally a dark circle, but now lit up, almost white. In quick succession came more urgent calls from the spacecraft. ‘‘We’ve got a fire in the cockpit!’’ yelled Chaffee, adding ‘‘let’s get out… we’re burning up’’ and finally uttering a brief scream. His words described the desperate bid by the astronauts to save themselves. Downstairs, on the first floor of Pad 34, technician Gary Propst, watching a monitor in the first seconds after the call, could clearly see Ed White, his arms raised above his head, fiddling with the hatch. Propst could not understand why the men did not simply blow the hatch, little realising that there was no way for them to do this.

Instead, White had to laboriously use a ratchet to release each of the six bolts spanning the circumference of the inner hatch. Years later, Dave Scott recalled that, during training, he and White had weightlifted the hatch over their heads whilst lying supine in their couches. Now, in the few seconds he had available before being overcome, White barely had chance to even begin loosening the first bolt. ft would have made little difference. As fire gorged its way through Apollo 1, the build-up of hot gases sealed the hatch shut with enormous force. No man on Earth could have opened the hatch under such circumstances. fn fact, even under the best conditions in the simulator, the inner hatch alone took 90 seconds to remove and none of the crew – even the super-fit White – had ever done it inside two minutes during training.

Subsequent investigations would determine that the fire had started somewhere beneath Grissom’s seat, perhaps in the vicinity of some unprotected and chafed wires and, once sparked in the pure oxygen atmosphere, fed itself hungrily and soon exploded into a raging inferno. Other readily-combustible products – Velcro pads, nylon netting, polyurethane padding and paperwork – added fuel to the flames. The astronauts themselves had even taken a block of styrofoam into the spacecraft to relieve the pressure on their backs throughout the test; it had exploded like a bomb in the pure-oxygen environment. “At such pressure, and bathed by pure oxygen,’’ wrote Grissom’s biographer Ray Boomhower, “a cigarette could be reduced to ashes in seconds and even metal could burn.’’

At length, pressures inside the cabin exceeded Apollo 1’s design limit and the spacecraft ruptured at 6:31:19 pm, filling Pad 34’s white room with thick black smoke and flames. By now, the pure oxygen had been guzzled by the fire and poisonous fumes had long since asphyxiated the three men. Outside, just a couple of metres away, North American’s pad leader Don Babbitt sprang from his desk and barked orders to lead technician Jim Greaves to get them out of the command module. ft was hopeless. The waves of heat and pressure were too intense and repeatedly drove the men back. “The smoke was extremely heavy,’’ Babbitt recalled. “ft appeared to me to be a heavy thick grey smoke, very billowing, but very thick.’’ fn fact, none of the pad staff could see far beyond their noses and had to physically run their hands over the outside of the boost cover to find holes into which they could insert tools to open the hatch.

The effects of the smoke were so bad that no fewer than 27 technicians were treated that night by Cape Kennedy’s dispensary for inhalation and Babbitt had to order Greaves out of the white room at one stage, lest he pass out. Eventually, with the assistance of more technicians and the arrival of firefighters, the hatch was opened and the would-be rescuers gazed at a hellish scene of devastation within. By the flickering glow of a flashlight, they could see nothing but burnt wiring, firefighter Jim Burch recounted, and it took a few seconds before they finally realised that the

The burnt-out remains of Apollo 1.

unreal calmness meant only one thing: Grissom, White and Chaffee were dead. It was 6:37 pm, five and a half minutes after the first report of fire. Choking over the phone to Slayton, Babbitt could not find words to describe what he saw.

Flight surgeon Fred Kelly, who arrived on the scene with Slayton shortly thereafter, was equally shocked. He observed that it would probably take hours to remove the three men from the spacecraft, because the intense heat had caused everything to melt and fuse together. Moreover, there remained the risk that the heat could accidentally set off the Saturn IB’s escape tower and the pad was cleared of all personnel. Not until the small hours of the following morning, 28 January, were the bodies removed. Grissom had detached his oxygen hose, probably in an effort to help White with the hatch, while Chaffee, in charge of communications, remained securely strapped into his seat. The autopsy team set to work immediately and found that none of the astronauts had suffered life-threatening burns; all had died from asphyxia when their oxygen hoses burned through and their space suits filled with poisonous carbon monoxide.

It was a devastating blow to the astronaut corps and to NASA as a whole. Deke Slayton described it as his “worst day’’ and Frank Borman, whom the agency appointed as its astronaut representative on the internal review board, admitted that he, Max Faget and Slayton “went out and got bombed’’ in the hours after the accident. “I’m not proud to say it,’’ continued the normally-teetotal Borman, “but… we ended up throwing glasses, like a scene out of an old World War One movie.’’ The wives of the three dead men – Betty Grissom, Pat White and Martha Chaffee – were also angry and would sue North American for its shoddy spacecraft. Each was awarded hundreds of thousands of dollars in compensation in 1972.

Three days after the accident, in flag-draped coffins, the bodies of the astronauts were met at Andrews Air Force Base in Washington, DC, by an honour guard and by representatives of NASA and President Lyndon Johnson’s administration. Grissom and Chaffee were laid to rest in Arlington National Cemetery, whilst White’s family insisted on his interment at the Military Academy at West Point.

The one saving grace of the tragedy was that it happened when it did, that it stopped NASA’s incessant ‘go’ fever in its tracks at a point from which there might be some recourse. If Grissom, White and Chaffee had been on their way to the Moon when such a fire erupted, it could conceivably have spelled the end of Apollo. Moreover, astronauts Tom Stafford, John Young and Gene Cernan, who were in Downey on the night of the disaster, preparing to back up Jim McDivitt’s Apollo 2 mission, felt the deaths of three men on the ground had probably saved six or more lives later on. By extension, the fact that investigators had the burnt-out spacecraft to examine meant that the cause could be pinpointed and rectified. A fire on the way to the Moon would have eliminated any chance of finding out what had happened.

“SMOOTH AS GLASS”

Meanwhile, preparation of the Gemini VIII flight hardware got underway in mid­August 1965, when the Agena-D target, tailnumbered ‘5003’, arrived at Cape Kennedy and, after final assembly, commenced pre-launch testing in October. Then, in early January of the following year, the two stages of the Titan II were delivered to Florida, mated and installed on Pad 19. Leak checks of the second stage engine on 7 February turned up small cracks in the thrust chamber manifold, a problem solved by rewelding, but by the 10th the rocket and Gemini VIII were fully mated and had undergone electrical compatibility tests.

Final preparation of the Agena-D occurred in tandem, the Atlas having been erected on Pad 14 at the Cape in early January. By this time, procedural and design changes – a result of Project Surefire – had been fully implemented. Two weeks later, the pencil-like Agena-D was mated to its docking adaptor and on 1 March was mounted atop the Atlas. The only problem in the last few days was an overfilling of its propellant tanks, which required the replacement of its regulator and relief valve and pushed the launch date from 15 March to the 16th.

In spite of the renewed vigour injected into the lunar effort by the spectacular Gemini VII/VI-A rendezvous late the previous year, morale in the astronaut corps suffered a devastating blow with the deaths of Elliot See and Charlie Bassett in the crash of their T-38 jet on 28 February 1966. Two days later, although their minds were almost wholly focused on their upcoming mission, Armstrong and Scott and their wives joined a huge crowd of mourners at Seabrook Methodist Church for See’s memorial service and at Webster Presbyterian Church for that of Bassett. A sense of foreboding pervaded the Gemini project. ‘‘People started… fearing we were in a run of bad luck,’’ wrote Dave Scott. ‘‘The memorial services in Houston… were sad, depressing affairs.’’ The following year, 1967, would be worse still.

Armstrong and Scott and their backups, Pete Conrad and Dick Gordon, had little time to waste on contemplation; a few days after the See and Bassett memorials, they were in quarantine at Cape Kennedy as launch drew closer. At 7:00 am on the morning of 16 March, Armstrong and Scott awoke and proceeded through the time – honoured ritual of a pre-flight physical and breakfast of filet mignon, eggs and toast with butter and jelly. Clear blue, cloud-speckled skies over the marshy Florida landscape promised a perfect opportunity for a launch. As they suited up, a watch belonging to aviation pioneer Jimmy Mattern – who had unsuccessfully attempted a round-the-world solo feat in 1933 – was strapped around the wrist of Armstrong’s suit. Not to be outdone, Scott, in his own nod towards aviation heritage, carried pieces of wood and cloth from an old Douglas World Cruiser, the New Orleans, which first flew around the world in 1924. Both artefacts had been borrowed from the museum at Wright-Patterson Air Force Base.

Shortly after insertion into his seat, a glitch with Scott’s parachute harness threatened a delay. One of the technicians discovered some epoxy in the catcher mechanism of the harness, but the efforts of Pete Conrad and pad fuerher Guenter Wendt finally got it unglued. “Pete… rushed around until he found a dentist’s toothpick with which to try and clear the connection out,’’ recalled Scott. “I remember looking back and seeing Pete sweating like mad digging this stuff out.’’ Conrad’s toothpick did the trick. The hatch was finally closed and Gemini VIII’s countdown proceeded.

One hundred minutes before launch, at 10:00 am, as their cabin atmosphere was steadily being purged with pure oxygen, the first stage of the mission got underway as the Atlas-Agena lifted-off successfully from Pad 14. Unlike Schirra and Stafford’s experience the previous October, the rocket performed near-flawlessly. Despite following a trajectory that was slightly low and to the south of its intended flight path, the Atlas’ sustainer engine compensated by pushing itself back on track, ultimately carrying the 6 m-long, pencil-like Agena-D into a circular orbit, 298 km high. Advised of the successful insertion, Armstrong told launch controllers that he and Scott were ready to go.

“Cradled in my contoured seat it felt almost as if I was being held in someone’s arms,’’ wrote Scott, comparing the sensation to taking a brand-new Ferrari out on the open road for the first time. The pure oxygen atmosphere introduced a cool cleanness and freshness to the cabin.

Liftoff occurred precisely on time at 11:41 am, conducted under the auspices, for the first time, of a flight director other than Chris Kraft. The distinctly English-accented, tweed-suited and pipe-smoking John Hodge, who had led one of three teams during the Gemini IV mission, was in charge. “The Titan was smooth,’’ continued Scott, describing a few shudders and evidence of pogo oscillations, but overall “a solid feeling, a sharp kick in the tail’’. His heart rate in the moments preceding liftoff peaked at 128 beats per minute; Armstrong’s touched 146. Post-flight aeromedical studies would see the difference simply as a ‘keying-up’ of Armstrong’s physical awareness, rather than an indicator of undue stress.

Launch and ascent aboard the Titan, Armstrong told James Hansen, was ‘‘very definite; you knew you were on your way when the rocket lit off. You could hear the thrust from the engines, at least at low altitudes, but the noise did not interfere with communications. . . The G levels got to be pretty high in the first stage of the Titan – something like 7 G’’. As the rocket headed higher into the rarefied atmosphere and

“SMOOTH AS GLASS”

Launch of an Atlas-Agena target vehicle.

its second stage took over, the two men saw bright red and yellow debris from a severed joining strap through their windows.

Riding the second stage, Scott wrote, was “smooth as glass”. Later, his first experience of weightlessness came when a small metal washer hovered in front of his eyes. As he released his checklist, he was amused to see it drift across the cabin. Nothing, however, could have prepared him for his first glimpse of Earth. Armstrong rolled Gemini VIII and the men beheld the deep blue of the Mediterranean, together with Italy and, in the distance, the outline of the Middle East and the Red Sea, laid out, map-like, before them. Although he had reached 63 km in his final X-15 flight and had seen the curvature of Earth, Armstrong, too, was quite unprepared for the view. In his autobiography, Scott remembered the difficulty other astronauts had when describing the sheer beauty and grandeur of the home planet; now he knew how they felt. Perhaps, someday, he thought, a poet or an artist would be blasted into the heavens. Maybe he or she could describe it better than a test pilot. “But I wanted to capture what I could,” he wrote. “I pulled out my camera and took my first photographs from space.”

Incredible as the view was, there was little time to dwell upon it. Within minutes, they established themselves in a 160 x 172 km orbit, trailing the Agena by 1,963 km. Scott’s spacewalk was looming in barely 24 hours’ time and, before that, they had the Agena-D rendezvous and docking to perform. A minor radiator problem did not prevent the astronauts from undertaking some sightseeing over the Pacific, a cloud – covered Hawaii, Baja California, the naval base in San Diego and the area around Edwards Air Force Base in the Mojave Desert, where Armstrong and Scott had both worked and studied years earlier.

Their first thruster firing came at 1:15 pm, a little more than 90 minutes after launch, which slightly lowered their apogee. A break for lunch took longer than anticipated, requiring them to Velcro-patch the food packages to Gemini VIII’s ‘ceiling’ whilst they executed a second burn on their second orbit to raise their perigee. When they came to eat lunch, it was hardly home cooking: Armstrong’s chicken and gravy casserole, despite having been rehydrated, was dry in places and the astronauts’ brownie cookies were stuck together and crumbly. ‘‘We had been running on adrenalin,’’ wrote Scott. ‘‘There had been no time for food – no thought of it either.’’ Nor was there time to spend worrying. A third manoeuvre, executed high above the Pacific Ocean at 2:27 pm, placed Gemini VIII into the same orbital plane as the Agena, albeit imprecisely.

‘‘A fundamental requirement of rendezvous,” Armstrong told James Hansen, ‘‘is to get your orbit into the same plane as the target’s orbit, because if you’re misaligned by even a few degrees, your spacecraft won’t have enough fuel to get to its rendezvous target. So the plan is to start off within just a few tenths of a degree of your target’s orbit. That is established by making your launch precisely on time, to put you in the same plane under the revolving Earth as is your target vehicle.’’

Armstrong then fired the aft OAMS thrusters, producing a horizontal velocity change of 8 m/sec, after which Capcom Jim Lovell requested that he add an extra 0.6 m/sec to his velocity. The adjustment, Armstrong said later, ‘‘was a pretty loose burn… without much preparation”, but the two astronauts quickly moved onto their next task: activating and testing the spacecraft’s rendezvous radar. Westinghouse, the company responsible for developing the device, had promised that it would be able to acquire its target at a distance of around 343 km.

They detected the Agena with their radar long before achieving a visual sighting, whilst still some 332 km away. A little under four hours into the mission, high above Madagascar, another burn adjusted their orbit. This prepared them with near­perfection to begin the ‘terminal’ phase of the rendezvous. An hour later, at 4:21 pm, during their third orbit, Scott reported his first visual sighting of the Agena as a speck in the distance, 140 km away, its rendezvous beacon flickering against the black sky. ‘‘Gradually, she became a sleek, silver tube,’’ he wrote, ‘‘a spectacular sight.’’

As both spacecraft drifted into orbital darkness, the Agena disappeared from view, although Armstrong and Scott were still able to discern its blinking acquisition lights. When the target was at the proper angle, ten degrees ‘above’ them, Armstrong recalibrated the platform of Gemini VUI’s inertial reference system for a translational manoeuvre. Next, he pitched the nose upwards 31.3 degrees and canted the spacecraft 16.8 degrees to the left, finally braking Gemini VIII by eyesight alone as Scott called out radar range and range rates. Several smaller thruster spurts brought them to a position just 46 m from the Agena, with no relative velocity and, after 30 minutes of inspections to ensure that the target had not been damaged during launch, they were given the go-ahead to execute the world’s first-ever docking manoeuvre.

Moving his spacecraft at barely 8 cm/sec, Armstrong gingerly pulsed towards the Agena, announcing the onset of the ‘station-keeping’ phase of the rendezvous at 5:40 pm. Years later, he told James Hansen that station-keeping at such close proximity to the target posed no major problems; conversations with Wally Schirra and Frank Borman had assured him that flying two vehicles close together was very easy to do once the correct position had been achieved.

For the next 25 minutes, he and Scott electronically checked the target’s systems, antennas and lights using radio command, before nudging their spacecraft closer; so close, in fact, that they could read a small, illuminated instrument panel above its docking cone. After receiving a go-ahead to dock from Keith Kundel, the capcom on the Rose Knot Victor communications ship, Armstrong pulsed Gemini VIII’s thrusters and achieved physical contact at 6:15 pm. An electric motor aboard the Agena retracted the docking cone, pulling the spacecraft’s nose about 60 cm into the target and connecting their electrical systems. By now close enough to read, the Agena’s display confirmed a green ‘rigid’ sign, indicating that both vehicles were mechanically and electronically mated.

‘‘Flight, we are docked,’’ Armstrong exulted, ‘‘and it’s really a smoothie. No noticeable oscillations at all.’’ Seconds later, as the realisation of what had been done finally set in, sheer pandemonium broke out in Houston. Armstrong and Scott’s ‘smoothie’ had cleared another hurdle on the road to the Moon.

That smoothie would rapidly give way to a far rockier road. . . one that would come close to claiming the two astronauts’ lives.

MOONSHIP

Project Apollo, which brought about the deaths of Grissom, White and Chaffee and which also enabled the steps of Neil Armstrong and Buzz Aldrin on the Moon, was born very soon after NASA’s own creation late in 1958. At that time, it was expected that exploration of the Solar System would be one arena in which the abilities of men, rather than machines, would be required. A fundamental obstacle, however, was the distinct absence of large boosters capable of fulfilling such roles and in mid-December of that year, newly installed Administrator Keith Glennan listened as Wernher von Braun, Ernst Stuhliner and Heinz Koelle presented the capabilities of existing hardware and stressed the need for a new ‘family’ of rockets. Landing men on the Moon, for the first time, was explicitly discussed as a long­term objective and, indeed, Koelle suggested a preliminary timeframe for achieving this as early as 1967.

Von Braun’s vision for the new family of rockets was that, first and foremost, their engines should be arranged in a ‘cluster’ formation, directly carrying an aviation concept into the field of spacegoing rocketry. The famed missile designer also discussed propellants and the idea of employing different combinations for different stages… then broached the subject of precisely how such enormous boosters could deliver a manned payload to the lunar surface. Von Braun had five methods in mind: one involving a ‘direct ascent’ from Earth to the Moon, the other four involving some sort of rendezvous and docking of vehicles in space. In whatever form the mission took, the rocket would need to be enormous, comprising, he said, ‘‘a seven-stage vehicle’’ weighing ‘‘no less than 6.1 million kg’’. Alternatively, he suggested flying a number of smaller rockets to rendezvous in Earth orbit and assemble a 200,000 kg lunar vehicle, which could then depart for the Moon. Aside from the immense practical problems of building and executing such a plan were the very real unknowns, Stuhlinger added, of how men and machines could operate in a weightless environment, with concerns of temperature, radiation, micrometeorites and corrosion an ever-present hazard.

Glennan’s focus at the time was, of course, Project Mercury, although in testimony before Congress early in 1959 he and his deputy, Hugh Dryden, admitted that there was ‘‘a good chance that within ten years’’ a circumnavigation of the Moon might be achieved, although not a landing, and that similar projects may be underway in connection with Venus or Mars. In support of NASA’s long-term aims, Glennan requested funding to begin developing the cornerstone for such epic ventures – the booster itself – and presented President Dwight Eisenhower with a report on four optional ‘national space vehicle programmes’: Vega, Centaur, Saturn and Nova. Although the first and last of these scarcely left the drawing board, the others would receive developmental funding and von Braun’s team, which had championed a rocket known as the Juno V, gained backing to develop it further under the new name ‘Saturn’.

In April 1959, Harry Goett, later to become director of NASA’s Goddard Space Flight Center in Greenbelt, Maryland, was called upon to lead a research steering committee for manned space exploration. The major conclusions of his panel were that, after Project Mercury had sent a man into orbit, the agency’s goals should encompass manoeuvring in space, establishing a long-term manned laboratory, conducting a lunar reconnaissance and landing and eventually surveying Mars or Venus. ‘‘A primary reason,’’ remarked Goett of the choice of the Moon as a major

target, “was the fact that it represented a truly end objective which was self-justifying and did not have to be supported on the basis that it led to a subsequent more useful end.”

Elsewhere, efforts to begin developing the Saturn were gathering pace. Its challenges, though, were both huge and staggering, with propellant weights alone for a direct-ascent rocket producing a vehicle of formidable scale; indeed, even the prospects for constructing a lunar spacecraft in Earth orbit would require more than a dozen ‘smaller’ launches and the added complexity of rendezvous, docking and assembly operations. At this early stage, the problems of being able to store cryogenics for long periods in space, to have a throttleable lunar-landing engine and takeoff engine with storable propellants and auxiliary power systems were first identified.

Unfortunately, midway through Dwight Eisenhower’s second term in office, and with the emphasis of his administration on balancing the budget ‘‘come hell or high water’’, it proved impossible for Glennan to formally commit NASA to a long-term lunar effort. Instead, small groups at the agency’s field centres began springing up, including one within the Space Task Group, which considered a second-generation manned vehicle capable of re-entering the atmosphere at speeds almost as great as those needed to escape Earth’s gravitational pull. ‘‘The group was clearly planning a lunar spacecraft,” wrote Courtney Brooks, James Grimwood and Loyd Swenson, and by the autumn of 1959 sketches of a lenticular re-entry vehicle had emerged and crystallised to such a point that its designers even applied for it to be patented.

Early January of the following year finally brought approval from Eisenhower for NASA to accelerate development of von Braun’s Saturn and offered the first hint of political support for manned space efforts beyond Project Mercury. Within weeks, Glennan’s request to Abe Silverstein, director of the Office of Space Flight Programs, to encourage advanced design teams at each NASA field centre and within the aerospace industry began to bear fruit: von Braun’s team proposed a Saturn-based lunar exploration design and J. R. Clark of Vought Astronautics offered a brochure entitled ‘A Manned Modular Multi-Purpose Space Vehicle’. At this point, of course, Project Mercury had yet to accomplish its first manned mission; however, regardless of their limited chances of receiving presidential or congressional approval, the proposals continued.

Other efforts focused on exactly how the spacecraft and other hardware could be delivered to the lunar surface in the most economical way. In May 1960, NASA’s Langley Research Center sponsored a two-day conference on rendezvous, with several techniques discussed, although it was recognised that they would be unlikely to bear fruit until the agency secured funding for a flight test programme.

It was at around this time that the decision was made over naming the spacecraft which would bring about the most audacious engineering and scientific triumph in the history of mankind. The name ‘Apollo’, formally conferred upon the programme on 28 July 1960 by Hugh Dry den, would honour the Greek god of music, prophecy, medicine, light and – perhaps above all – progress. ‘‘I thought the image of the god Apollo riding his chariot across the Sun,’’ wrote Abe Silverstein, who had consulted a book on mythology to come up with the name, ‘‘gave the best representation of the grand scale of the proposed programme.’’

The scope of Apollo, Bob Gilruth and others revealed to more than 1,300 governmental, scientific and industry attendees at a planning session in August, was for a series of Earth-orbital and circumlunar expeditions as a prelude for the first manned landing on the Moon. Guidelines for the design of the spacecraft would be fourfold: it would need to be compatible with the Saturn booster under development by von Braun’s team, it had to be able to support a crew of three men for a period of up to a fortnight and it needed to encompass the lunar or Earth-orbital needs of the project, perhaps in conjunction with a long-term space station. By the end of October, three $250,000 contracts were awarded to teams led by Convair, General Electric and the Martin Company for initial studies.

In spite of this apparent brightening of the lunar project’s chances, Glennan himself remained unconvinced that Apollo was ready to move beyond the feasibility stage and felt a final decision would have to await the arrival of the new president in January 1961. By this point, Glennan was estimating Apollo to cost around $15 billion and felt that the Kennedy administration needed to spell out, clearly, and with no ambiguity, its precise reasons for pursuing the lunar goal, be they for international prestige or scientific advancement. At around the same time, Hugh Dryden and Bob Seamans directed George Low to head a Manned Lunar Landing Task Group, detailed to draft plans for a Moon programme, utilising either direct – ascent or rendezvous, within cost and schedule guidelines, for use in budget presentations before Congress. When Low submitted his report in early Lebruary, he assured Seamans that no major technological barriers stood in the way and that, assuming continued funding of both the Saturn and Apollo, a manned lunar landing should be achievable between 1968-70.

Moreover, Low’s committee was considerably more optimistic than Glennan in terms of cost estimates: they envisaged spending to peak around 1966 and total some seven billion dollars, reasoning that by that time the Saturn and larger Nova-type boosters would have been built and an Earth-circling space station would probably be in existence. It stressed, however, that manned landings would require a launch vehicle capable of lifting between 27,200 and 36,300 kg of payload; the existing conceptual design, dubbed the ‘Saturn C-2’, could boost no more than 8,000 kg towards the Moon. Low’s group advised either that several C-2s needed to be refuelled in space or an entirely new and more powerful booster awaited creation. Both approaches seemed realistic, the committee concluded, with Earth-orbital rendezvous probably the quickest option, yet still requiring the technologies and techniques to refuel in space.

Of pivotal importance in the subsequent direction of Apollo was the new president, John Litzgerald Kennedy, who had already appointed a group before his inauguration to assess the perceived American-Soviet ‘missile gap’ and investigate ways in which the United States could pull ahead technologically. The group was headed by Jerome Wiesner of the Massachusetts Institute of Technology – later to become Kennedy’s science advisor – and it advocated, among other points, that NASA’s goals needed to be both redefined and sharpened. Another key figure, and long-time ally of NASA, was Vice-President Lyndon Johnson, who pushed strongly to appoint James Webb, a man with immense experience in government, industry and public service, to lead NASA. On 30 January 1961, Webb’s appointment as the agency’s second administrator was authorised by Kennedy.

It was Webb who would guide NASA through the genesis of Apollo; indeed, his departure from the agency would come only days before the project’s first manned launch in October 1968. His importance to America’s space heritage and the respect in which he continues to be held will be recognised, just a few years from now, by the launch of the multi-billion-dollar James Webb Space Telescope (JWST), successor to Hubble. Yet Webb’s background was hardly scientific or in any way related to space exploration: a lawyer by profession, he directed the Bureau of the Budget and served as Undersecretary of State for the Truman administration, but throughout the Sixties he would prove NASA’s staunchest and most fierce champion.

Also championing the agency’s corner was President Kennedy himself, who, only weeks before Yuri Gagarin’s flight, raised its budget by $125 million above the $1.1 billion appropriations cap recommended by Eisenhower. Much of this increase was funnelled into the Saturn C-2 development effort and, specifically, its giant F-1 engine. Built by Rocketdyne, the F-1 – fed by a refined form of kerosene, known as ‘Rocket Propellant-1’ (RP-1), together with liquid oxygen – remains the most powerful single-nozzled liquid engine ever used in service. Although it experienced severe teething troubles during its development, particularly ‘combustion instability’, it would prove impeccably reliable and the cornerstone to a lunar landing capability.

By this time, Convair, General Electric and the Martin Company had submitted their initial responses to NASA, none of which overly impressed the agency’s auditors; indeed, recounted Max Faget, all three had stuck rigidly with the same shape as the Mercury capsule. Some theoreticians had already predicted that a Mercury-type design would be unsuitable for Apollo’s greater re-entry speeds and Space Task Group chief design assistant Caldwell Johnson had begun investigating the advantages of a conical, blunt-bodied command module.

Early in May 1961, after more adjustments and rework, the contractors offered their final proposals to NASA. Convair envisaged a three-component Apollo system, its command module nestled within a large ‘mission module’. Notably, it would return to Earth by means of glidesail parachute and develop techniques of rendezvous, docking, artificial gravity, manoeuvrability and eventual lunar landings. General Electric offered a semi-ballistic blunt-bodied re-entry vehicle, with an innovative cocoon-like wrapping to provide secondary pressure protection in case of cabin leaks or micrometeoroid punctures. Martin, lastly, proposed the most ambitious design of all. Conical in shape, its Apollo was remarkably similar to the design ultimately adopted, although it featured a pressurised shell of semi – monocoque aluminium alloy coated with a composite heat shield of superalloy and a charring ablator. Its three-man crew would sit in an unusual arrangement, with two abreast and the third behind, in a set of couches which could rotate to better absorb the G loads of re-entry and enable better egress.

All three contractors spent significantly more than the $250,000 assigned by NASA, with Martin’s study topping three million dollars, requiring the work of 300 engineers and specialists and taking six months to complete. In their seminal work on the development of Project Apollo, Brooks, Grimwood and Swenson pointed out that, had times been less fortunate, NASA may have been obliged to spend months evaluating the contractors’ reports before making a decision. However, it was at this time that Yuri Gagarin rocketed into orbit and John Kennedy pressed Lyndon Johnson to find out how the United States could beat the Russians in space. On 25 May 1961, before a joint session of Congress, he made the lunar goal official… and public.

In the wake of Kennedy’s speech, one of the key areas into which the increased funding would be channelled was a new booster idea called ‘Nova’; this was considered crucial to achieving a lunar landing by the direct-ascent method. At this stage, although NASA was ‘‘studying’’ orbital rendezvous as an alternative to direct – ascent, Hugh Dryden explained that ‘‘we do not believe… that we could rely on [it]’’. More money and increased urgency for Apollo was not necessarily a good thing: both Webb and Dryden felt that decisions over direct-ascent or orbital rendezvous and liquid or solid propellants would have been better made two years further down the line.

Nonetheless, rendezvous as an option was steadily coming to the fore, with a realisation that it could provide a more attractive alternative to the need for enormous and unwieldy boosters, instead allowing NASA to use two or three advanced Saturns with engines that were already under development. Although Earth-orbital rendezvous was considered safer, a lunar-orbit option would require less propellant and could be done with just one of von Braun’s uprated Saturn ‘C-3’ rockets.

The Apollo spacecraft which would fly missions to the Moon was also taking shape. Max Faget, the lead designer of the Space Task Group, set the diameter of its base at 4.3 m and rounded its edges to fit the Saturn for a series of test flights. These rounded edges also simplified the design of an ablative heat shield which would be wrapped around the entire command module. Encapsulating the spacecraft in this way provided additional protection against space radiation, although on the downside it entailed a weight penalty. Others, including George Low, saw merits in both blunt-bodied and lifting-body configurations and suggested that both should be developed in tandem. Most within the Space Task Group, however, felt that a blunt body was the best option.

Notwithstanding these issues, in August 1961 NASA awarded its first Apollo contract to the Massachusetts Institute of Technology, directing it to develop a guidance, navigation and control system for the lunar spacecraft. Two months later, five aerospace giants vied to be Apollo’s prime contractor, with the Martin Company ranked highest in terms of technical approach and a very close second in technical qualification and business management. In second place was North American Aviation, whom the NASA selection board recommended as the most desirable alternative. On 29 November 1961, word quickly leaked out to Martin that its scores had won the contest to build Apollo, but proved premature; the following day, it was announced by Webb, Dryden and Seamans that North American would be the prime contractor, in light, it seemed, of their long-term association with NASA and NACA and their spaceflight experience. The choice of North American, whose fees were also 30 per cent lower than Martin, would in many minds return to haunt NASA in years to come.

Rumour quickly abounded that it was politics, and not technical competency, which had won North American the mammoth contract. Astronaut Wally Schirra would recount that he felt the decision was made because companies in California had yet to receive their fair share of the space business, while others pointed to the company’s lobbyist Fred Black, who had developed a close relationship with Capitol Hill insider Bobby Gene Baker, a protege of Vice-President Lyndon Johnson.

As North American and NASA hammered out their contractual details, the nature of Apollo’s launch vehicle remained unclear, as, indeed, was its means of reaching the Moon. It was likely that the production of large boosters capable of accomplishing a direct-ascent mission would take far longer than the development of smaller vehicles. The attractions of rendezvous were also becoming clearer as a means of meeting Kennedy’s end-of-the-decade deadline. At around this time, Bob Gilruth wrote that “rendezvous schemes may be used as a crutch to achieve early planned dates for launch vehicle availability and to avoid the difficulty of developing a reliable Nova-class launch vehicle’’.

As the debate over the launch vehicle continued, it was recognised that, in whatever form it took, it would be enormous and would demand a correspondingly enormous launch complex. Under consideration were Merritt Island, north of Cape Canaveral, together with Mayaguana in the Bahamas, Christmas Island, Hawaii, White Sands in New Mexico and others. Only White Sands and Merritt Island proved sufficiently economically competitive, flexible and safe to undergo further study. The final choice: a 323 km2 area of land on Merritt Island for a site later to become known (after the assassination of President John Kennedy) as the Kennedy Space Center. One of the most iconic structures to be built here in the mid-Sixties, and associated forever with the lunar effort, was the gigantic Vehicle (originally ‘Vertical’) Assembly Building (VAB), used to erect and test the Saturn rockets. Standing 160 m tall, 218 m long and 158 m wide, it covered 32,400 m2 and to this day remains the world’s largest single-story building.

Elsewhere, a site near Michoud in Louisiana was picked for the Chrysler Corporation and Boeing to assemble the first stages of the Saturn C-1 and subsequent variants. In October 1961, NASA purchased 54 km2 in south-west Mississippi and obtained easement rights over another 518 km2 in Mississippi and Louisiana for a static test-firing site for the large booster, prompting around a hundred families, including the entire community of Gainsville, to sell up and relocate. It was around the same time that the decision to move the Space Task Group – now superseded by the Manned Spacecraft Center – from Virginia to Houston, Texas, was made.

On the morning of 27 October 1961, shortly after 10:06 am, the maiden mission in support of Apollo got underway with the test of the Saturn 1 (originally C-1) rocket from Pad 34 at Cape Canaveral. Although the vehicle was laden with dummy upper stages, filled with water, its performance was satisfactory, but its 590,000 kg of thrust was woefully insufficient to send men to the Moon and back. Still, it marked the first of ten Saturn 1s launched, which, by the time of its last flight in July 1965, had carried a ‘boilerplate’ command and service module into orbit. Most engineers envisaged the lunargoing Saturn would need at least four or even five F-1 engines in its first stage. This would permit an Earth-orbit or lunar-orbit rendezvous mode to deliver a payload to the Moon’s surface. Despite continuing interest in a large, direct-ascent Nova, employing as many as eight F-1s, the decision was taken on 21 December to proceed with a rocket known as the Saturn C-5 (later the Saturn V), capable of supporting both Earth-orbital and lunar-orbital rendezvous missions.

However, direct ascent was still considered by many as the safest and most natural means of travelling to the Moon, sidestepping the dangers of finding and docking with other vehicles in space. Yet procedures for exactly how a lander might be brought onto the lunar surface remained sketchy, with some suggesting the bug-like spacecraft touching down vertically on deployable legs or horizontally on skids. An Air Force-funded study, begun in 1958 and called ‘Lunex’, had already addressed a direct-ascent method of reaching the Moon. However, Wernher von Braun doubted it was possible to build a rocket large enough to accomplish such a mission and favoured rendezvous with smaller vehicles. Before coming to NASA, von Braun’s team had proposed a mission known as ‘Project Horizon’, which justified the need for a lunar base for military, political and, lastly, scientific purposes. He felt that only Saturn was powerful enough to complete such a mission and one of his conditions upon joining NASA was that its development should continue.

Against this backdrop came the appearance of the lunar-orbital rendezvous plan, whereby a craft would descend to the Moon’s surface and, after completing its mission, return to rendezvous with a ‘mother ship’. The landing crew would then transfer to the orbiting mother ship and return to Earth. Since 1959, in fact, this idea had been recognised as the best technique to reduce the total weight of the spacecraft. Many within NASA, however, were terrified by the prospect of attempting rendezvous so far from home. Proponents, on the other hand, considered it relatively simple, with no concerns about weather or air friction, lower fuel requirements and no need for a monster Nova rocket. ft was NASA engineer John Houbolt who finally convinced Bob Seamans to place it on an equal footing with direct ascent and Earth-orbital rendezvous when a decision came to be made. By July 1962, the decision had been made: lunar-orbital rendezvous would be adopted, employing a separate lander in addition to the command ship.

At the same time, the first steps to actually design the lander got underway, with early plans ranging from short-stay missions involving one man for a few hours to seven-day expeditions with crews of two. One design took the form of an open, Buck Rogers-like ‘scooter’ with landing legs, which the fully-suited astronaut would manoeuvre onto the surface. As these plans crystallised, the paucity of knowledge of the lunar surface material, and the effect of exhaust gases on its rocks and dust, made it imperative that astronauts could ‘hover’, brake their spacecraft and select an appropriate landing spot.

North American, which had already been awarded the contract to build the command and service modules, strongly opposed the lunar-orbital rendezvous mode, partly because it wanted its spacecraft to perform the landing. (fndeed, in August 1962, cartoons adorned its factory walls, depicting a somewhat disgruntled Man in the Moon looking suspiciously at an orbiting command and service module and declaring ‘‘Don’t bug me, man!’’) With this in mind, North American made a strong bid to build the lander, which NASA rejected on the basis that the company already had its hands full with the development of the main spacecraft. By September 1962, 11 companies had submitted proposals to build the lander and in November the Grumman Aircraft Engineering Corporation of Bethpage, New York, was chosen for the $388 million contract. Although each bidder was judged technically and managerially capable, Grumman had spacious design and manufacturing areas, together with clean-room facilities to assemble and test the lander.

The decision to proceed with lunar-orbital rendezvous eliminated the requirement for the Apollo command module to land on the Moon, but created a new problem: the need for a form of docking apparatus by which it could link up with Grumman’s lander. The need was quickly identified for a series of Earth-orbital missions to demonstrate and qualify the command module’s systems before committing them to lunar sorties; the result was the Block 1 and 2 variants, the second of which provided the docking hardware and means of getting to the Moon. By mid-1963, North American had begun work on an extendable probe atop the command module, which would fit into a dish-shaped drogue on the lunar lander.

As the design of the command module moved through Block 1 and 2 variants, so the lunar module itself was changing into its final form: a two-part, spider-like ‘bug’ which would deliver astronauts to the Moon’s surface and back into orbit. Its four­legged descent stage would be equipped with the world’s first-ever throttleable rocket engine, whilst the ascent stage, housing the pressurised cabin, would have a fixed – thrust engine to boost the crew back into lunar orbit. The organic appearance of the lunar module produced something which Brooks, Grimwood and Swenson described as ‘‘embodying no concessions to aesthetic appeal. . . ungainly looking, if not downright ugly’’. Operating within Earth’s atmosphere, obviously, would be unnecessary and aerodynamic streamlining was ignored by the Grumman designers. However, when the time came for the ascent stage to liftoff from the lunar surface, its exhaust in the confined space of the inter-stage structures – ‘fire-in-the-hole’ – could produce untoward effects, perhaps tipping the vehicle over. Clearly, many problems remained to be solved.

Shape-wise, the ascent stage was originally spherical, much like that of a helicopter, with four large windows for the crew to see forward and ‘down’. This design was ultimately discarded when it became clear that the windows would need extremely thick panes and strengthening of the surrounding structure. Two smaller windows were chosen instead, but the need for visibility remained very real, eliminating the spherical cabin design in favour of a cylindrical one with a flat forward bulkhead cut away at various planar angles. The windows became small, flat, triangular panels, canted ‘downwards’ so that the crew would have the best possible view of the landing site.

Changing from a spherical to a cylindrical cabin, though, meant that Grumman’s engineers could not easily weld the structure. By May 1964, they had decided to weld areas of critical structural loads, but rivets would be employed where this was impractical. The interior of the 4,930 kg ascent stage cabin, with a volume of 60 m3, made it the largest American spacecraft yet built and NASA pressed Grumman to make its instruments as similar to those in the command module as possible. As it evolved, the astronauts became an integral part of it, with Pete Conrad working on the design perhaps more so than anyone else. He was instrumental in implementing electroluminescent lighting inside the lunar module, as well as the command module, reducing weight and power demands.

Another crucial change in the design of the lunar module was the removal of seats, which were seen as too heavy and restrictive in view of the fact that the astronauts would be clad in bulky space suits. Bar stools and metal cage-like structures were considered, but the brevity of the lunar module’s flight and moderate G loads eventually rendered them totally unnecessary. Moreover, standing astronauts would have a better view through the windows and the eliminated worry about knee room meant that the cabin could be reduced in size. Instead of seats, restraints would be added to hold the astronauts in place and prevent them from being jostled around during landing.

The hatch, through which the astronauts would exit and re-enter from the lunar surface, was changed from circular to square to make it easier for their pressurised suits and backpacks to fit through. At the base of the 10,334 kg descent stage were five legs, later reduced to four as part of a weight-versus-strength trade-off, and 91 cm footpads with frangible probes to detect surface impact. Keeping the lander’s weight down was of pivotal importance, to such an extent that NASA paid Grumman $20,000 for every kilogram they could shave off. Even the weight of the astronauts helped determine which of them would fly the lunar module and which would not.

Inside the third stage of the Saturn V launch vehicle, the lander’s legs would be folded against the structure of the descent stage and extended in space. In addition to its ascent and descent engines, the lunar module possessed 16 small attitude-control thrusters, clustered in quads, pointing upwards, downwards and sideways around the ascent stage for increased manoeuvrability. The ascent engine, built by Bell, was a key component which simply had to work to get the astronauts away from the lunar surface; as a result, it was the least complicated device, with a pressure-fed fuel system employing hypergolic propellants. The descent engine was more challenging, since it had to be throttleable: Rocketdyne, its builder, used helium injection into the propellant flow to decrease thrust while maintaining the same flow rate.

As the command, service and lunar modules took shape, the launch vehicles for the Earth-orbital (Saturn 1B) and lunar (Saturn V) missions also approached completion. The two-stage Saturn 1B – Gus Grissom and Wally Schirra’s “big maumoo’’ – underwent its first test on 26 February 1966 and also marked the first ‘real’ flight of a ‘production’ Apollo command and service module. The rocket’s S-IB stage had arrived at Cape Kennedy in mid-August of the previous year, followed by the S-IVB a month later. By the end of October, the rocket’s instrument unit and the command and service module for the mission, designated ‘Apollo-Saturn 201’ (AS – 201), were in Florida. After numerous delays, including lower-than-allowable pressures in the S-IVB, the flight got underway at 11:12 am. The S-IB carried the Saturn to an altitude of 57 km, whereupon the S-IVB took over and boosted AS-201 to an altitude of 425 km.

After raising its own apogee to 488 km, the command and service module’s SPS engine was ignited to accelerate its return to Earth. Splashdown came at 11:49 am, half an hour after launch, and the undamaged spacecraft was hauled aboard the recovery vessel Boxer. Despite problems, AS-201 proved that the Apollo spacecraft was structurally sound and that its heat shield could survive a high-speed re-entry. However, its SPS had not performed as well as expected; firing, but only operating correctly for about 80 seconds, after which its pressure fell by 30 per cent due to helium ingestion into its oxidiser chamber. Managers, obviously, did not want such an event to occur during a return from the Moon. The SPS problem had to be rectified. Further, the effects of microgravity on the propellants in the S-IVB, which would be needed to perform the translunar injection burn, needed to be better understood.

Consequently, a decision was taken to reverse the plan of unmanned Saturn 1B launches for the remainder of the year. The six-hour AS-203 mission, not planned to carry a command and service module, was shifted ahead of AS-202 and launched on 5 July. It satisfactorily demonstrated that the S-IVB’s single J-2 engine could indeed restart in space and that the propellants behaved exactly as predited. Seven weeks later came AS-202, during which the SPS was fired four times without incident, demonstrating its quick-restart capabilities, and the heat shield was tested. Its 90- minute mission cleared the way for Apollo 1, still internally dubbed ‘AS-204’, at the end of the year.

“HOW MANY BOYS, LBJ?”

Days after Gemini IV hit the waves of the Atlantic, the United States Military Command in South Vietnam announced that its troops would shortly begin fighting alongside South Vietnamese forces against the pro-communist North. It was the beginning of a long, bloody and infamous phase of American history that would see a markedly different United States by the end of the decade compared to that which John Kennedy had inherited in 1961.

Yet Kennedy himself was at least partly to blame for the steady escalation in the conflict in south-east Asia, as was his presidential successor, Lyndon Baines Johnson (LBJ). In fact, one of the issues Kennedy had faced during the I960 election was a perceived ‘missile gap’ between America and the Soviets and in his inaugural address he had promised to ‘‘pay any price, bear any burden, meet any hardship, support any friend, oppose any foe, in order to assure the survival and success of liberty’’. hollowing the crisis of the Bay of Pigs and the erection of the Berlin Wall, however, Kennedy feared that if his administration did nothing to halt the incessant advance of communism from North into South Vietnam, the United States would lose all credibility with its allies.

His initial move was to increase American troop numbers and, although he felt that President Ngo Dinh Diem must ultimately defeat the communist insurgents, the poor organisation, corruption and incompetent leadership of the South Vietnamese military made this an unlikely prospect. A key facet in the effort to isolate the populace from the communists was the Strategic Hamlet Programme, established in 1961, yet even its provisions of improved education and healthcare did little to prevent its infiltration by guerrilla factions. Moreover, the South Vietnamese peasantry resented being uprooted from their ancestral villages. As a result, by 1963, the project had all but collapsed.

That summer, some Washington policymakers were predicting that Diem’s inability to subdue the communists might force him to make a deal with North Vietnamese premier Ho Chi Minh. His main concern seemed to be the need to fend off military coups against himself and, according to Attorney-General Bobby Kennedy, “Diem… was difficult to reason with’’. Early suggestions to encourage a coup and forcibly remove Diem from power were ultimately discarded, for fear of the potential destabilising effects of such an action, and it was decided instead that his younger brother, the hated chief of the secret police, Ngo Dinh Nhu, be removed instead. Only weeks before Kennedy’s assassination, the CIA advised anti-Diem generals that the United States would support his removal from office. On 2 November, during a failed attempt to escape, Diem and Nhu were captured and summarily executed in the back of an armoured personnel carrier.

Kennedy, who had not approved such an act, was visibly shocked and in the wake of his own assassination three weeks later the South Vietnamese situation bounced from one military junta to another, all of which were regarded as nothing more than puppets of the United States. Their existence served only to encourage North Vietnam to consider the leaders of the South as slaves of colonialism. Indeed, in the wake of the Diem and Nhu murders, the number of American ‘military advisors’ multiplied to 16,300 to cope with rising guerrilla violence. Fears that the pro-communist Vietcong retained de facto control of the South Vietnamese countryside prompted the United States to adopt a different policy of pacification and ‘‘winning over hearts and minds’’.

In the aftermath of Kennedy’s assassination, Johnson asserted his own support for continued military operations in South Vietnam, before the end of 1963 pledging $500 million in aid to Saigon. In August of the following year, the destroyer Maddox, on an intelligence mission along the North Vietnamese coastline, fired upon several torpedo boats in the Gulf of Tonkin. Two days later, an alleged attack on the Maddox and another destroyer, the Turner Joy, prompted the Americans to initiate air strikes, marking their first large-scale military involvement in the conflict. Although the second incident was later discovered to be an error, it led Congress to pass the Gulf of Tonkin Resolution, giving Johnson the authority to assist any south-east Asian nation under threat of communist aggression. Over the years, the Gulf of Tonkin ‘error’ has prompted some observers to argue that Johnson deliberately misled the American populace to gain support for greater involvement in Vietnam. Others refute this, but suggest that Robert McNamara and the Pentagon were in the mood to retaliate and presented their ‘evidence’ of the attacks to support their case.

In February 1965, an attack on a Marine barracks in Pleiku led to the ignition of Operations Flaming Dart and Rolling Thunder, a pair of vigorous aerial bombing campaigns to force North Vietnam to terminate its support for the Vietcong. The campaigns would last for three years, depositing, by November 1968, over a billion kilograms of missiles, rockets and bombs into North Vietnam, along the Ho Chi Minh Trail in Laos and Cambodia and onto Vietcong installations in South Vietnam. Ultimately, it failed.

A month after the Pleiku attack, 3,500 Marines were deployed to South Vietnam to help defend the United States’ air bases, effectively beginning the ground war, with public opinion overwhelmingly supporting the move as part of a global conflict against communism. By the end of the year, troop numbers had swelled to 200,000. It was a commitment which many would regret as the Sixties wore on and the death toll rose, with the phrase “Hey, hey, LBJ! How many boys did you kill today?’’ chanted by protesters on the steps of the Pentagon in late 1967.

By the second half of the Sixties, with almost half a million American troops in Vietnam, many of the astronauts were developing itchy feet to return to active military service. “We were uncomfortable wearing the hero image,’’ wrote Gene Cernan, “while our buddies were bleeding, being captured and dying in a real shooting war for which we had been trained.’’ Some of them, indeed, approached Deke Slayton with a view to taking leave of absence from NASA, polishing up their carrier landing qualifications and returning to the front line… only to be awakened to the harsh reality. They were free to leave, if they wished, Slayton told them, but he offered no guarantees of a job when they returned.

“The Pentagon hammered in the final nail,’’ continued Cernan. “We could return to active duty if we wanted to, and even fly, but never – ever – would we be allowed into combat.’’ The negative propaganda impact of the Vietcong capturing American astronauts in a combat zone was too unpalatable for the Johnson administration to bear. “Vietnam,” wrote Cernan, “would not be our war.’’

“TUMBLING END-OVER-END”

Gemini VIII was not aided by a distinct lack of available tracking stations across its flight path, which resulted in some very spotty communications with Mission Control. “We could communicate with Houston,” wrote Scott, “for only three five – minute periods every 90 minutes as we passed over secondary tracking stations.” Two ship-based stations were aboard the Rose Knot Victor and the Coastal Sentry Quebec, plus a land-based site in Hawaii. Shortly before one such loss of contact, at 6:35 pm, Armstrong and Scott were advised by Jim Lovell that, if problems arose whilst docked, they should deactivate the Agena and take control with the Gemini. “Just send in the command 400 to turn it off,” Lovell told them, “and take control with the spacecraft.” For now, however, the only problems seemed to be difficulties verifying that the Agena was receiving uplinked commands and a glitch with its velocity meter, used to specify the magnitude of a burn by its main engine.

Twenty-seven minutes after docking, Scott commanded the Agena to turn them 90 degrees to the right and Armstrong reported that the manoeuvre had “gone quite well”. His call came seconds before Gemini VIII passed out of radio range of the land-based Tananarive station in the Malagasy Republic. Working alone, the astronauts transmitted an electronic signal to start the Agena’s tape recorder. Shortly thereafter, their attitude indicators showed them to be in an unexpected, almost imperceptible, 30-degree roll. “Neil,” called Scott, “we’re in a bank’’.

Perhaps, they wondered, the Agena’s attitude controls were playing up or its software load was wrong. Since Gemini VIII’s OAMS was now switched off and both men could see the Agena’s thrusters firing, they reasoned that the target’s controls must be at fault. They could not have known at the time that one of their own OAMS thrusters – the No. 8 unit – had short-circuited and stuck in its ‘on’ position. Unaware, Scott promptly cut off the Agena’s thrusters, whilst Armstrong pulsed the OAMS in a bid to stop the roll and bring the combined spacecraft under control. For a while, his efforts succeeded.

After four tense minutes, the docked spacecraft slowed and steadied itself. Then, as Armstrong worked to reorient them into their correct horizontal position, the unwanted motions began again. . . much faster than before and, wrote Scott, ‘‘on all three axes’’. Perplexed, the men jiggled the control switches of the Agena, then of the Gemini, off and on, in a fruitless effort to isolate the problem. At around this time, Scott noticed that Gemini VIII’s own attitude propellant had dropped to just 30 per cent, clearly indicative of a problem with their own spacecraft. ‘‘It was clear,’’ Scott related, ‘‘we had to disengage from the Agena, and quickly.’’

Undocking presented its own problems, not least of which was the very real risk that the two rapidly rotating vehicles could impact one another. However, Scott, demonstrating an intuitive test pilot’s awareness of the importance of recording all pertinant data, pre-set the Agena’s recording devices such that ground controllers would still be able to remotely command it. ‘‘I knew that, once we undocked, the rocket would be dead,’’ he wrote. ‘‘No one would ever know what the problem had been or how to fix it.’’ Scott’s prompt action saved the Agena and preserved it for subsequent investigations and tasks.

Still out of radio contact with the ground, Armstrong moved onto the next step of the flight rules and undocked from the Agena. He then fired a long burst of Gemini VIII’s translational thrusters to pull away, only to discover that the spacecraft, now free, began to spin more wildly, in roll, pitch and yaw axes. It was now much worse than before, because the stuck-on No. 8 thruster was no longer turning the entire combination, only the Gemini. High above south-east Asia, they came within range of the Coastal Sentry Quebec, which received Scott’s urgent radio transmission at 6:58 pm: “We have serious problems here… we’re tumbling end-over-end. We’re disengaged from the Agena.’’

Aboard the ship, Capcom Jim Fucci acknowledged the call and enquired as to the nature of the problem. Both men were relieved to hear Fucci’s voice. “He was an old NASA hand,’’ wrote Scott, “very experienced.” Quickly, yet with characteristic calmness, Armstrong reported that he and Scott were “in a roll and we can’t seem to turn anything off. . . continuously increasing in a left roll’’. Fucci duly passed the

“TUMBLING END-OVER-END”

Gemini VIII during rendezvous activities with the Agena.

report over to Houston: Gemini VIII was suffering “pretty violent oscillations”. The three-way conversation with Mission Control meant that it was some seconds before Flight Director John Hodge picked up all of the details; Fucci having to repeat that “he’s in a roll and he can’t stop it”.

Armstrong quickly threw circuit breakers to cut electrical power and hence the flow of propellant to the attitude thrusters, including the No. 8 unit. However, with no friction or counterfiring thruster to stop it, the spinning continued. At its worst, it reached 60 revolutions per minute. Everything in the cabin – checklists, flight plans, procedures charts – were hurled around by the centrifugal force. The unfiltered Sun flashed in the astronauts’ windows with alarming regularity, said Scott, “like a strobe light hitting us in the face’’.

Such a rotation rate placed the men at serious risk of physically blacking out and, indeed, both had difficulties reading their instruments properly. “It was rather like the feeling you get as a kid when you twist a jungle rope round and round and then hang on it as it spins and unfurls,’’ wrote Scott. “In space, it was not a good feeling.’’ Physician Chuck Berry would later note that the astronauts experienced two conditions brought on by the rapid rotation: a complete loss of orientation caused by the effects on their inner ears (the ‘coriolis effect’), coupled with ‘nystagmus’, an involuntary rhythmic motion of the eyes.

The incessant rotation and the depletion of Gemini VIII’s attitude propellant had already alerted Armstrong and Scott to a problem with their own spacecraft, but at the time they did not know that the short circuit in the OAMS had caused the No. 8 thruster to become stuck ‘on’ and caused the rapid drop in fuel. Even had they known, there would have been no time to ponder it. Armstrong’s responsibility as the command pilot was to ensure the safety and success of the mission. Scott’s spacewalk, docked activities with the Agena and most of the experiments were now off the agenda; the safe return of the crew was paramount.

Armstrong decided that his only available course of action was to use Gemini VIII’s 16 re-entry controls. It was easier said than done. The re-entry control switch was in the most awkward position imaginable, directly above Armstrong’s head, and, worse, was on a panel with around a dozen toggles. ‘‘With our vision beginning to blur,’’ wrote Scott, ‘‘locating the right switch was not simple.’’ Fortunately, both men had carried their years of test-piloting experience into the astronaut business and intuitively knew every switch, literally, with their eyes closed. ‘‘Neil knew exactly where that switch was without having to see it,’’ Scott continued, but admitted ‘‘reaching above his head. . . while at the same time grappling with the hand controller… was an extraordinary feat.’’

The effort to reduce the spacecraft’s rates to zero with the re-entry controls, though ultimately successful, consumed 75 per cent of the fuel. ‘‘We are regaining control of the spacecraft slowly,’’ Armstrong reported, as the spinning stopped within 30 seconds. The flight, however, was over. Mission rules decreed that the re­entry controls, once activated, would require an immediate return to Earth at the next available opportunity. Just ten hours into its three-day mission, Gemini VIII was on its way home.

WITCH HUNT

The fire in the AS-204 spacecraft on 27 January 1967 left plenty of blame to go around and both North American, whose workmanship was seen as shoddy, and NASA, who had overseen them and given their seal of approval, were savaged by the media, by the public and by lawmakers alike. The media, indeed, were making up their own stories. On 10 February, for example, Time magazine cited the New York Times as having quoted an unidentified official who claimed that Grissom, White and Chaffee had screamed repeatedly for help in those frantic seconds. Their bodies, the official added, had been incinerated. . .

Fearing that Congress could pull the plug on Apollo with immediate effect, the agency set to work on the night of the disaster on its own internal review, with an eight-man panel headed by Langley Research Center director Floyd Thompson. Although Olin Teague, chair of the House Space Subcommittee, was keen for NASA to complete its work, others within the Senate were impatient and called for a hearing on 27 February. There, Administrator Jim Webb was verbally grilled, with representatives condemning ‘‘the level of incompetence and carelessness” as ‘‘just unimaginable”. Recriminations took an uglier turn when Senator Walter Mondale probed Webb for details of something called ‘The Phillips Report’.

Apollo’s programme manager, a retired Air Force general named Sam Phillips, had strongly criticised North American’s performance as prime contractor for over a year. He considered their relationship with NASA to be quarrelsome and disagreeable and had established a ‘tiger team’ to inspect the situation. This had

Seated before a Senate hearing, NASA’s senior management were verbally grilled and NASA’s “carelessness” and “incompetence” were particularly attacked. From left to right are Bob Seamans, Jim Webb, George Mueller and Sam Phillips.

left him with serious concerns, so much so that on 16 December 1965 he wrote a scathing memo to North American chairman Lee Atwood, placed the company on notice to improve and told George Mueller, NASA’s associate administrator for the Office of Manned Space Flight, that he had “lost confidence” in the prime contractor. Now, in the spring of 1967, Jim Webb revealed that he had never been made privy to the contents of Phillips’ report.

Others, including North American inspector Thomas Baron, had since 1965 condemned the level of poor workmanship they saw at Cape Kennedy, together with infractions of cleanliness and safety rules. Although Baron’s judgements were refuted by North American in its congressional testimony, they cannot have helped to quieten those who were looking for blame. Some, including the writer Erik Bergaust in his 1968 book ‘Murder on Pad 34’, even implied that NASA had blood on its hands for racing recklessly with the decade and killing the three men in the process.

Against this backdrop of public and media fury, the Thompson board worked for ten weeks, assisted by 1,500 technicians, and traced all possible sources of fire in Apollo’s 30 km of electrical wiring and even re-enacted the blaze in a command module mockup. Additionally, Spacecraft 014, the Block 1 vehicle originally assigned to Wally Schirra’s Apollo 2 mission, was shipped from Downey to Cape Kennedy for systematic dismantling and inspection alongside the burnt-out Spacecraft 012. Cabin pressures, the investigators found, had soared from the normal test pressure of 1.15 bars, slightly above sea-level equivalent, to 2.0 bars, rupturing the spacecraft’s hull, but it was Bureau of Mines expert Robert van Dolah who revealed the damning truth: an escape hatch, capable of being opened in a couple of seconds, might have saved the astronauts. Thompson’s report was published on 9 April and ran to 3,300 pages. It found no definitive cause for the fire, but suspected an unexplained arc on wiring beneath Grissom’s left footrest, which spurted to another object and ignited the 100 per cent oxygen atmosphere.

The report cited ‘‘deficiencies in command module design, workmanship and quality control’’, including uncertified and highly-flammable materials in the cabin, as having contributed to the tragedy. Additionally, it revealed that many safety checks simply were not done, nor was there enough fire-suppression equipment at Pad 34. ‘‘It was,’’ wrote Deke Slayton, ‘‘about as scathing a document as you’d ever see from a government agency towards itself.’’

Days later, Thompson and others found themselves testifying before the House and quickly discovered that even pro-Apollo congressmen were fiercely unsympa­thetic. Some lawmakers even went so far as to suggest reviewing the business of selecting contractors for the lunar effort. At one stage, responding to a question from Congressman John Davis of Georgia, North American’s John McCarthy raised the possibility that Grissom himself might have inadvertently started the fire by kicking a batch of loose wires. Although Slayton admitted that McCarthy’s comment was only raised in response to a question, he wrote that ‘‘it really pissed me off… because there were no grounds for the story – it was pure speculation, not to mention physically impossible’’.

The effect of the fire elsewhere in the space agency was equally dramatic. Bob Gilruth, who had become a virtual father figure to many of the astronauts, broke down in tears upon learning of the tragedy. In his autobiography, Wally Schirra recalled taking him out on for a spin on his Cal 25 sailboat a few months after the accident and, whilst manning the tiller, Gilruth fell asleep. “Maybe it was the first chance he’d had to relax, to realise he had to push ahead and forget the tragedy,” Schirra wrote. “Gilruth was carrying a tremendous load.’’ So too was Joe Shea, the man who might have been inside the command module, sitting in precisely the spot where the fire started that terrible evening. He took the fire very badly, shifting into overdrive in an impossible personal crusade to solve Apollo’s problems… and, in doing so, drove himself to the brink of a breakdown. Eventually, he was moved to NASA Headquarters, then left to work for Raytheon. Years later, Shea would wonder if he could have snuffed out the fire… and convinced himself, with 70 per cent certainty, that he could have successfully smothered it.

It was the straight-talking Frank Borman who summed up what should happen in testimony on 17 April. “Let’s stop the witch hunt,’’ he told Congress, “and get on with it.’’ Getting on, though, would involve more than a year and $75 million-worth of changes to turn Apollo into a very different machine to that in which Grissom, White and Chaffee had died. Its cabin would now be pressurised with a mixture of 60 per cent oxygen and 40 per cent nitrogen, then steadily replaced with pure oxygen at partial pressure after launch as the nitrogen leaked out. No major structural reworking of the command module would be necessary. All flammable materials were to be removed and, crucially, a new 32 kg single-piece hatch was implemented, which opened outwards and could be sprung in just five seconds. Its mechanism, assisted by a cylinder of compressed nitrogen gas, could be opened with a little finger.

Elsewhere, aluminium plumbing, which melted at 580°C, was replaced by stainless steel, and coolant pipelines which could release flammable glycol when ruptured were ‘armour-plated’ with high-strength epoxy. Wire bundles were encased in protective metal panels and nylon netting and plastic containers were replaced by fire-retardant materials such as Teflon. Intricate ‘Velcro maps’ were created to limit the presence of this useful, but highly flammable, material and identify exactly where every piece of it would be located in the command module’s cabin. Paperwork was kept to a minimum, to such an extent that the crews were barred from taking reading materials with them. ‘‘No books or magazines,’’ wrote Wally Schirra. ‘‘Nor could we take anything made of paper to play with, such as cards or puzzles. We would find boredom a serious problem as we progressed through ten days in orbit.’’

The space suits to be worn by the astronauts had their nylon outer coatings replaced by beta cloth – an advanced fibreglass material produced by Owens – Corning Fibreglass Corporation – and supported by 14 layers of fire-resistant material. ‘‘We’re paying a price for safety,’’ Apollo 7 flight director Glynn Lunney told Time magazine. ‘‘The suits are bulkier, the fibreglass itches like hell and the seat belts are difficult to cinch down because they are so stiff, but you are seeing a spacecraft several hundred per cent improved.’’ Further, an emergency venting system capable of reducing the cabin pressure in seconds provided an extra safeguard to snuff out fires. Overall, the changes increased Apollo’s weight by 1,750 kg and placed it just beneath the Saturn V’s total lifting capacity for lunar missions. As a

result, parachutes were enlarged to permit safer splashdowns at greater weights, some redundant systems were eliminated and lead ballast was removed.

By extension, of course, the disaster which had befallen the command module could also afflict Grumman’s lunar module and increased fervour was placed on reviewing its materials, too. Nylon-based items were replaced by beta cloth and ‘booties’ were installed over circuit breakers to lessen the risk of electrical shorts. This work on the lunar module – the machine which would actually set men on the Moon – refocused attention on the key question: would John Kennedy’s dream ever be realised, within the decade, or at all? At the beginning of 1967, NASA had spent $23 billion on Project Apollo and many now questioned the need for America to go there. The continuing threat of the Soviet Union provided one reason: Leonid Brezhnev’s increasingly regressive and repressive regime had, only a year before, consigned writers Yuli Daniel and Andrei Sinyavsky to hard labour for penning satirical, anti-Soviet texts. In some minds, it harked back to far darker times under Stalin.

When physicist Edward Teller was asked by Congress what he expected men to find on the Moon, he replied: ‘‘The Russians!’’ Even now, the sense of fear was as strong as ever. For their part, the Russians, mysteriously, had been conspicuously absent from the manned spaceflight business for almost two years by the time of the Apollo 1 fire, but their ambitions in Earth orbit were ready for a new resurgence. The death of Sergei Korolev and the appearance of a successor, Vasili Mishin, had pushed the Soviet Union’s new spacecraft – Soyuz (‘Union’) – further and further behind schedule. Now, three months after the deaths of Grissom, White and Chaffee, it was ready to go. Or was it?

WISEGUY

“That’s a vagina,’’ quipped Charles ‘Pete’ Conrad Jr. “Definitely a vagina.’’ The psychiatrist noted his response without a word, perhaps realising, perhaps not, that he was the victim of yet another wisecrack from the gap-toothed, balding Navy lieutenant. Yet, despite his comments about each of the Rorschach cards shown to him, Conrad was not entirely obsessive about the female genitalia. He had actually been tipped-off the night before by another astronaut candidate, Al Shepard: what the NASA psychiatrists were really looking for was male virility. ‘‘I got the dope on the psych test,’’ Shepard had assured him. ‘‘No matter what it looks like, make sure you see something sexual.’’ So Conrad did.

His key concern, though, that spring in 1959, had been the impact that this crazy ‘Project Mercury’ idea might have on his career. Instead of logging hours in the Navy’s new F-4 Phantom fighter, he spent a week at the Lovelace Clinic in Albuquerque, New Mexico, much of his time focused on the provision of stool,

semen and blood samples and the collection of 24-hour bagfuls of urine. On the evening before a major stomach X-ray, told not to drink alcohol after midnight, Conrad had sat up until 11:57 pm draining a bottle with Shepard and another naval aviator called Wally Schirra to loosen themselves up for the next day. Conrad doubted that Lovelace’s invasive tests had anything remotely to do with spaceflight: the physicians, he told Shepard, seemed far more interested in “what’s up our ass’’ than in their flying abilities. Shepard had warned him to be careful – to give the right answers to questions and to remember that Lovelace’s staff were watching their every move.

In spite of his frustration, Conrad persevered. He followed Shepard’s advice, saw the female anatomy in every Rorschach card, deadpanned to a psychologist that one blank card was upside down, pedalled a stationary bicycle for hours, sat in a hot room for an age, then dunked his feet into ice-cold water and argued with one of the physicians that he considered it pointless to have electricity zapped into his hand through a needle. However, all this torture, Conrad felt, would at least give him the opportunity to lay his entire naval career on the line for just one chance to fly something even faster: to ride a rocket, outside Earth’s atmosphere, “at a hell of a lot more Machs than anything he was flying right now’’. Flying higher and faster, and pushing his own boundaries, had been the story of Conrad’s life.

Born in Philadelphia, Pennsylvania, on 2 June 1930, the offspring of a wealthy family which made its fortune in real estate and investment banking, Conrad’s father insisted that he be named ‘Charles Jr’ – “no middle name’’ – although his strong – willed mother, Frances, felt that this tradition of Charleses should be broken. Frances liked the name ‘Peter’, wrote Nancy Conrad in her 2005 biography of her late husband, and although it never became his official middle name, Charles Conrad Jr would become known as ‘Peter’ or ‘Pete’ for the rest of his life. His fascination with anything mechanical reared its head at the age of four, when he found the ignition key to his father’s Chrysler and reversed it off the drive. Later, in his teens, he worked summers at Paoli Airfield, mowing lawns, sweeping and doing odd jobs for free flights. Aged 16, he even repaired a small aircraft single-handedly. Conrad was an engineer and tinkerer at heart.

Education-wise, he would partly follow in his father’s footsteps: the private Haverford School, from which he was expelled, then the Darrow School in New York, where Conrad’s dyslexia was identified and where he shone. Although his father intended him to attend Yale University, he actually enrolled in 1949 at Princeton, with a Reserve Officers Training Corps (ROTC) scholarship from the Navy to pay for his studies in aeronautical engineering. Graduation in 1953 brought him not only his bachelor’s degree, but also a pilot’s licence with an instrument rating, marriage (to Jane DuBose) and entrance into naval service.

He breezed through flight training, earning the callsign ‘Squarewave’ as a carrier pilot. In ‘Rocketman’, his widow wrote that Al Teddeo, executive officer of Fighter Squadron VF-43 at Naval Air Station Jacksonville, Florida, had his doubts when he first met the young, seemingly-wet-behind-the-ears ensign one day in 1955. Those doubts were soon laid to rest when Teddeo discovered that Conrad could handle with ease any manoeuvre asked of him. Tactical runs, strafing runs, spin-recovery tests; Ensign Conrad did it all. “Hell, we refuelled three times till I just had to get back to my desk,” Teddeo recalled years later. “It was like telling a kid at the fair that it was time to go home.”

Next came gunnery training at El Centro, California, and transition from jet trainers to the F-9 Cougar fighter, before reporting to Pax River in 1958 to qualify as a test pilot. Later that same year, he received, along with over a hundred others, classified instructions to attend a briefing in Washington, DC. Conrad was told to check into the Rice Hotel under the cover name of ‘Max Peck’. Only when he got there did he find that another 35 ‘Max Pecks’ were also there – including an old naval buddy, Jim Lovell. Neither Conrad nor Lovell would make the final cut for the Mercury selection, but their day would come three years later.

Whereas Lovell was cast aside for a minor liver ailment, however, Conrad’s cause for failure proved a little ironic. “Unsuitable for long-duration flight,’’ read the explanatory note. He had, it seemed, shown a little too much cockiness and independence during testing; characteristics which were at loggerheads with the panel’s notion of a good, all-rounded, level-headed astronaut. Six years after reading those words, Conrad and his Gemini V command pilot, Gordo Cooper, would rocket into space and set a new record… for long-duration spaceflight!

PACIFIC RETURN

On the ground, the television networks – which had cancelled their showings of ‘Batman’ and, ironically, ‘Lost in Space’ – were deluged with complaints from viewers as attention turned to a dramatic recovery effort. Original plans called for Gemini VIII to land in the Atlantic and be picked up by the aircraft carrier Boxer; however, the earlier-than-expected return called instead for a splashdown in the western Pacific during their seventh orbit.

The timing was strict. Gemini VIII’s flight path had precessed so far westwards that it would be another full day before Armstrong and Scott could reach a location from which they could be easily recovered. Consequently, a naval destroyer named the Leonard F. Mason, based off the coast of Vietnam, was directed to intercept the new splashdown point, 800 km east of Okinawa. It would be the only Gemini splashdown in the Pacific, in a landing zone designated ‘Dash 3’. ‘‘I looked it up in our manuals,’’ wrote Scott. ‘‘Dash 3 was a secondary landing zone in the South China Sea. It was over 6,000 miles away from our primary landing site.’’

It was far from ideal. By this time, John Hodge’s ‘blue’ flight control team had been at their consoles for 11 hours and a second (‘white’) team, headed by Gene Kranz, reported for duty to supervise the end of the mission. Kranz’s team had more experience in recovery procedures than that of Hodge and, had Gemini VIII run to its intended three-day length, he would have overseen re-entry and splashdown anyway. It made sense, therefore, for Kranz to take the helm.

The news of an impending return was met with grim resignation by Armstrong and Scott, who ran through their pre-retrofire checklists with the capcoms at the Coastal Sentry Quebec, Rose Knot Victor and Hawaiian tracking stations. Unlike Gemini V, which had been nursed through a lengthy mission, despite problems, the situation in which Armstrong and Scott found themselves was compounded by a dangerously-low propellant load. By this time, having tested each of the OAMS thrusters in a now-stable Gemini VIII, Armstrong had identified the glitch with the No. 8 unit, which Scott later described as not exhibiting ‘‘a consistent, linear problem… it was really screwed up’’. In fact, the thruster had been off when it should have been on, and vice-versa, on several occasions. The cause, however, would have to wait for the post-flight investigation.

Loading the re-entry flight program into Gemini VIII’s 4,000-word-memory computer was difficult, particularly as it was already overloaded from the rendezvous with the Agena. This required Scott to erase the rendezvous and docking programs, then feed the re-entry data into the computer by means of a keypad and an on-board device known as an auxiliary tape memory unit. As he worked to punch in a series of nine lines of seven-digit numbers, Scott was relieved that the unflappable Jim Fucci, aboard the Coastal Sentry Quebec, was there to watch his every move. ‘‘He read off those numbers as if he was talking about taking a stroll in the park,’’ Scott wrote. ‘‘I entered them quickly so that I could transmit them back to verify with him before we lost contact again.’’

Gemini VIII’s retrorockets duly ignited at 9:45 pm, whilst out of radio contact, high above a remote part of south-central Africa. Worse, retrofire was conducted during orbital darkness, giving Armstrong and Scott no horizon by which to judge alignment. Minutes later, over the Himalayas, the spacecraft entered the tenuous upper atmosphere and as it continued to descend through the steadily thickening air, Scott reported that he could see nothing but a pinkish-orange glow through his window… then haze and, finally, minutes before splashdown, the glint of water! Ten hours and 41 minutes after leaving Cape Kennedy, at 10:22 pm, the spacecraft hit the Pacific with a harsh thump and Scott yelled “Landing Safe!”

Throughout all this – during the launch, rendezvous, docking, crisis with and without the Agena, re-entry and splashdown – Jimmy Mattern’s watch, tightly strapped around Armstrong’s wrist, continued to tick faithfully…

Despite having suffered severe space sickness and, now, seasickness as the spacecraft’s windows rhythmically rolled and pitched with each wave, the astronauts swiftly proceeded through their post-splashdown checklist, shutting down electrical systems, placing switches and valves into their correct positions and activating their high-frequency communications antenna. Only now did Armstrong and Scott regret not taking Mission Control’s advice to swallow meclizine motion sickness tablets before re-entry. “When Mission Control told us about three-foot waves,’’ Scott wrote, “they had forgotten to mention the 20-foot swells!’’

Scott called the search-and-rescue team from Naha Air Base in Okinawa by their callsign ‘Naha Rescue One’, but was met with silence on the radio. Both men were hot in their suits, particularly Scott, whose ensemble had extra layers to provide radiation protection on his spacewalk. Fumes from the ablated heat shield, too, left them nauseous. Within half an hour, a C-54 aircraft, flown by Air Force pilot Les Schneider, which had spotted Gemini VIH’s descent and splashdown, arrived on the scene. Its crew visually checked the spacecraft, marked its landing co-ordinates and dropped three pararescue swimmers and an emergency liferaft. For the Naha Rescue One team, which was more accustomed to missions in Vietnam, Laos and Cambodia in those war-charged times, 16 March 1966 was a distinctly different and highly memorable day.

Notwithstanding the rough swells, the pararescue swimmers, themselves queasy, affixed a flotation collar to the spacecraft, then signalled the C-54 with a ‘thumbs-up’ that Armstrong and Scott were alive and well. This was duly radioed to other aircraft in the area, to the Leonard F. Mason, then to Hawaii, to NASA’s Goddard Space Flight Center and finally to Mission Control in Houston, from where public affairs officer Paul Haney announced the news to an anxious world. Meanwhile, the encounter between the antiquated C-54 and the state-of-the-art Gemini was, said Neil Armstrong, ‘‘the most unusual rendezvous in aviation history’’.

Three hours after splashdown, in the small hours of 17 March, the two astronauts and their spacecraft were safely aboard the Mason. The rough seas, though, had made the hoisting of Gemini VIII difficult, to such an extent that it kept crashing against the side of the destroyer, denting its nose at one point. The Mason’s crew, wrote Scott, had initially been less than happy about being given the task of recovering Gemini VIII. They had just completed a seven-week tour in Vietnam and been given a brief spell of liberty in Okinawa. However, their spirits rose as the realisation set in that the astronauts were safe. In spite of their tiredness and the

PACIFIC RETURN

Scott and Armstrong, surrounded by recovery swimmers after performing the Gemini project’s first – and only – splashdown in the Pacific.

 

effects of nausea, Armstrong and Scott managed smiles and greetings for the crew and were found to be healthy, suffering from minimal dehydration.

They were, however, shaken by what had actually come close to disaster… as, indeed, had many within NASA. Deputy Administrator Bob Seamans had been advised of the crisis over the telephone whilst at the reception to the prestigious Robert H. Goddard Memorial Dinner and swore that he would never again be caught in such a position during the critical phase of a future mission.

At the same time, publicly, NASA was reluctant to over-emphasise the near­disaster, particularly if it wanted continued funding for a Moon landing by 1970. When Life magazine proposed titling its Gemini VIII article as ‘Our Wild Ride in Space by Neil and Dave’, its editor-in-chief received a firm request from Armstrong to change it to something less melodramatic. Ultimately, bound by an ongoing contract, the magazine agreed and would publish watered-down headlines for Gemini VIII and subsequent missions.

In spite of the troubles, President Lyndon Johnson reassured the American public that his administration remained firmly committed to John Kennedy’s goal of bootprints on the Moon before the end of the decade. Some have argued over the years that Armstrong’s coolness was pivotal in his selection to command Apollo 11, although some isolated individuals within the astronaut office speculated that his status as a civilian test pilot had contributed to the failure.

Indeed, Walt Cunningham, later to fly Apollo 7, would criticise what he saw as flaws in both astronauts’ performance, while Tom Stafford felt that the decision to undock from the Agena was a flawed one. Gene Kranz, on the other hand, perceived the crisis as the result of a broader training failure – malfunction procedures did not cover the problems encountered whilst the Gemini and Agena were docked – and both Frank Borman and Wally Schirra praised Armstrong and Scott’s actions as having prevented disaster. Indeed, without their safe return and the knowledge of what had happened, an erroneous assumption that the Agena was to blame could have diseased the final days of Gemini and made it very difficult for Apollo, with its emphasis on rendezvous and docking, to proceed. ‘‘It could have been a showstopper,’’ admitted Dave Scott.

Gene Cernan, though, rationalised the critics’ thinking. ‘‘Screwing up was not acceptable in our hypercompetitive fraternity,’’ he told James Hansen. ‘‘Nobody got a free ride when criticism was remotely possible. Nobody.’’ Still, Gemini VIII did little damage to either man’s career. Definitive testament came two weeks after the flight, when the Gemini VIII Mission Evaluation Team “positively ruled out’’ any errors on the astronauts’ part and, indeed, Bob Gilruth himself praised them for their ‘‘remarkable piloting skill’’. Scott was promoted to lieutenant-colonel and assigned a seat on an Apollo crew within days, while Armstrong received the backup command slot for Gemini XI. Still, the quiet civilian was demoralised by what he saw as only a partial success.

Had he been ‘‘smarter’’, Armstrong said later, he might have figured out the problems earlier, perhaps saving Scott’s EVA and some of the mission’s other objectives. Many of Gemini VIII’s experiments – the zodiacal light photography task, the growth of frogs’ eggs, the synoptic terrain studies, the nuclear emulsions

PACIFIC RETURN

Armstrong (left) and Scott with crewmen aboard the recovery ship Mason.

and the cloud spectrophotography – were left incomplete and some have speculated over the years that, had Scott’s EVA been underway when the spinning started, he may have seen the burst from the stuck-on No. 8 thruster and warned Armstrong to shut off its propellant.

However, others considered it fortuitous that Scott’s EVA had never come to pass. It “had seemed terribly complex and dangerous,’’ wrote Mike Collins. The need for Scott to get outside, manoeuvre himself to Gemini VIII’s adaptor section and worry about swapping connectors and keeping track of tethers was, in Collins’ mind, too risky at such an early stage. “My own EVA scheme on Gemini X was far from ideal,’’ he wrote, “in that I had to stuff everything into an already crowded cockpit, but at least I could make nearly all my preparations inside the pressurised cocoon… Not so with Dave’s complicated gear.’’

Other naysayers have added that, during the uncontrollable spinning, Scott may have been whirled around so violently on his tether as to have hit the side of Gemini VIII, almost certainly producing fatal injuries…

OUTER SPACE TREATY

On the evening that the Apollo 1 crew lost their lives, the astronaut office in Houston was unusually quiet. At one point, only Al Bean was on duty and it was he who received the first word from Cape Kennedy of the fire. Several other astronauts were at Downey, California, running through simulations and practice for their missions… and a select delegation was at the White House in Washington, DC. There, veteran astronauts Scott Carpenter, Gordo Cooper, Jim Lovell, Neil Armstrong and Dick Gordon witnessed the signing by President Johnson of a document popularly called ‘The Outer Space Treaty’. Four decades later, the document has around a hundred signatories and a further two dozen who are partway through their ratification of it.

Officially, it is known as ‘The Treaty on Principles Governing the Activities of States in the Exploration and Use of Outer Space, Including The Moon and Other Celestial Bodies’. Essentially, the document forms the basis for the earliest international space law and on the very day that Grissom, White and Chaffee died, it was opened for signing by the United States, Great Britain and the Soviet Union. Its 17 articles decree that signatories will refrain from the placement of nuclear weapons or weapons of mass destruction into Earth orbit, onto the Moon or onto any other celestial body. The treaty explicitly states that the Moon and other celestial bodies are to be used for peaceful purposes and forbids weapons-testing and military exercises or implacements on them. Moreover, it denies signatories the right to ‘claim’ a celestial resource, such as the Moon, as its own and declares all to be “province of mankind’’. It also assures the safe and cordial return of any astronauts or cosmonauts who make an unexpected landing within the borders of another nation.

The astronauts liked to call it the ‘‘non-staking-a-claim treaty’’ and as the afternoon wore into evening, they mingled with guests at the event, including ambassadors from the Soviet Union (Anatoli Dobrynin), Great Britain (Patrick Dean) and Austria (Kurt Walheim, later Secretary-General of the United Nations). In his biography of Armstrong, James Hansen noted the astronaut’s recollection that the event ended at 6:45 pm and that, with the exception of Carpenter, the NASA delegation returned to the Georgetown Inn on Wisconsin Avenue. When they got to their rooms, they were greeted by flashing red lights on their answer machines. Something terrible had happened in Florida. A difficult year lay ahead.

“EIGHT DAYS OR BUST”

Although Gemini V, the first to carry and utilise fuel cells for electrical power, had long been planned to fly for seven or even eight days, the success of its predecessor and the performance of Jim McDivitt and Ed White had emboldened NASA to move up their estimates for the first lunar landing from 1970 to 1969 and, perhaps, said Joe Shea, as early as mid-1968. Both Gemini IV astronauts would remain very much part of the unfolding action: White was named within weeks to the backup command slot for Gemini VII, an assignment rapidly followed by the coveted senior pilot’s seat on the maiden Apollo voyage. McDivitt, too, would go on to great things: commanding Apollo 9, a complex engineering and rendezvous flight to pave the way for the first Moon landing. He would even be offered, but would refuse, the chance to walk on the lunar surface himself.

First, though, came the adulation. After an initial Houston reception, they headed for Chicago, where a million people greeted them and showered them in tickertape along State Street and Michigan Avenue. This was followed, in Washington, DC, by another parade down Pennsylvania Avenue to the Capitol, receptions in the Senate, meetings with foreign diplomats and even a free trip to Paris to upstage the appearance of Yuri Gagarin and a mockup of Vostok 1 at 1965’s Air Show. It is unknown to see such scenes as tickertape parades for astronauts today and, perhaps, the only ones in the foreseeable future may be for the men and women who return to the Moon or become the first to tread the blood-red plains of Mars.

In the Sixties, however, every mission was heroic. Moreover, despite the appalling workload and the inevitable strain the astronaut business placed on marriages and families, every man who left Earth’s atmosphere was a fully-fledged hero. Not for nothing did Gerry and Sylvia Anderson name their five Thunderbird heroes after five of the heroes of the Mercury Seven: Alan, Virgil, John, Scott and Gordon. For one of those heroes, Gordo Cooper, and his rookie pilot, Pete Conrad, the reality in the build-up to their mission was one of exhausting 16-hour workdays, plus weekends, and a tight schedule to launch on 1 August 1965, eight weeks after McDivitt and White splashed down. Cooper and Conrad and their backups, Neil Armstrong and Elliot See, had only been training since 8 February, giving them less than six months to prepare for the longest mission yet tried. “We realised they needed more time,” wrote Deke Slayton. “I went to see George Mueller to ask him for help and he delayed the launch by two weeks.”

Despite the pressure, Cooper and Conrad found time to give some thought to names for their spacecraft, even though NASA had officially barred them from doing so. Due to its pioneering nature, the two men wanted to call Gemini V ‘The Conestoga’, after one of the broad-wheeled covered wagons used during the United States’ push westwards in the 18th and 19th centuries. Their crew patch, in turn, would depict one such wagon, emblazoned with the legend ‘Eight Days or Bust’. This was quickly vetoed by senior managers, who felt it suggested a flight of less than eight days would constitute a failure, and Conrad’s alternative idea – ‘Lady Bird’ – was similarly nixed because it happened to be the nickname of the then-First Lady, wife of President Johnson. Its possible misinterpretation as an insult could provoke unwelcome controversy which NASA could ill-afford. The astronauts, however, would not be put off and Cooper pleaded successfully with Jim Webb to approve the Conestoga-wagon patch, although the administrator greatly disliked the idea. The duality of the word ‘bust’ as denoting both a lack of success and the female breasts did not help matters, either. . .

Preparations for Gemini V had already seen Conrad gain, then lose, the chance to make a spacewalk. According to a January 1964 plan, the Gemini IV pilot would depressurise the cabin, open the hatch and stand on his seat, after which an actual ‘egress’ would be performed on Gemini V (Conrad’s mission), a transfer to the back of the spacecraft and retrieval of data packages on Gemini VI and work with the Agena-D target vehicle on subsequent flights. Following the Voskhod 2 success, however, plans for a full egress were accelerated and granted to Ed White. The result: instead of ‘Eight Days or Bust’, Gemini V would come to be described by Cooper and Conrad as ‘Eight Days in a Garbage Can’; they would simply ‘exist’ for much of their time aloft, to demonstrate that human beings could survive for at least the minimum amount of time needed to get to the Moon and back. (The maximum timespan for a lunar mission, some 14 days, would be an unwelcome endurance slog earmarked for the Gemini VII crew.)

Yet the Conestoga mission did have its share of interesting gadgets: it would be the first Gemini to run on fuel cells, would carry the first production rendezvous radar and was scheduled to include exercises with a long-awaited Rendezvous Evaluation Pod (REP). Originally, it was also intended to fly the newer, longer-life OAMS thrusters, although these were ready ahead of schedule and incorporated into Gemini IV. Only weeks after Cooper, Conrad, Armstrong and See began training, on 1 April 1965 fabrication of the Gemini V capsule was completed by McDonnell,

“EIGHT DAYS OR BUST”

A tired and heavily-bearded Conrad (left) and Cooper aboard the recovery ship after the flight.

 

Подпись: 270 Pushing the Envelope

tested throughout May in the altitude chamber and finally delivered to Cape Kennedy on 19 June. Elsewhere, GLV-5 – the Titan booster assigned to launch the mission – was finished in Baltimore, accepted by the Air Force and its two stages were in Florida before the end of May. Installation on Pad 19 followed on 7 June, the day of McDivitt and White’s splashdown, and Gemini V was mounted atop the Titan U on 7 July. Five days later, the last chance for an EVA on the mission and, indeed, on Geminis VI and VII, was rejected by NASA Headquarters. There seemed little point in repeating what White had already done and, further, Cooper and Conrad, not wishing to be encumbered by their space suits for eight days, had campaigned vigorously for greater comfort in orbit by asking to wear helmets, goggles and oxygen masks. The launch of Gemini V was scheduled for 19 August.

It would be a false start. Thunderstorms ominously approached the Cape, rainfall was copious and a lightning strike caused the spacecraft’s computer to quiver. The latter, provided by IBM, had caused concern on Gemini IV and, this time around, had been fitted with a manual bypass switch to ensure that the pilots would not be left helpless again. The attempt was scrubbed with barely ten minutes remaining on the countdown clock and efforts to recycle for another try on 21 August got underway. On this second attempt, no problems were encountered. Aboard Gemini V, Cooper turned to Conrad. “You ready, rookie?’’ Conrad, white as a sheet, replied that he was nervous. Surely the decorated test pilot who had flown every supersonic jet the Navy owned wasn’t scared? Conrad milked the silence in the cabin for a few seconds, then burst out laughing. “Gotcha!” he said with his trademark toothy grin. “Light this son-of-a-bitch and let’s go for a ride!’’ And ride they did. At 8:59:59 am, Cooper and Conrad were on their way.

Ascent was problematic when noticeable pogo effects in the booster jarred the men for 13 seconds, but smoothed out when the second stage ignited and were minimal for the remainder of the climb. Six minutes after launch, as office workers across America snoozed away their Saturday morning, Gemini V perfectly entered a 163-349 km orbit. Nancy Conrad wrote that her late husband compared the instant of liftoff to “a bomb going off under him, then a shake, rattle and roll like a ’55 Buick blasting down a bumpy gravel road – louder than hell’’.

Hitting orbit made Cooper the first man to chalk up two Earth-circling missions. (Gus Grissom, of course, had piloted a suborbital flight on Liberty Bell 7, before commanding the orbital Gemini 3.) However, Gemini V would shortly encounter problems. The flight plan called for the deployment of the 34.5 kg REP, nicknamed ‘The Little Rascal’, from the spacecraft’s adaptor section, after which Cooper would execute a rendezvous test, homing in on its radar beacon and flashing lights. Before the REP could even be released, as Gemini V neared the end of its first orbit, Conrad reported, matter-of-factly, that the pressure in the fuel cells was dropping rapidly from its normal 58.6-bar level. An oxygen supply heater element, it seemed, had failed. Nonetheless, as they passed over Africa on their second orbit, Cooper yawed the spacecraft 90 degrees to the right and, at 11:07 am, explosive charges ejected the REP at a velocity of some 1.5 m/sec. Next, the flight plan called for Gemini V to manoeuvre to a point 10 km below and 22.5 km behind the REP, although much of this work was subsequently abandoned. However, Chris Kraft’s ground team was becoming increasingly concerned as the fuel cell pressures continued to decline and when a pressure of 12.4 bars was reached this was insufficient to operate the radar, radio and computer. Kraft had little option but to tell the astronauts to cancel their activities with the pod.

It seemed likely that a return to Earth would be effected and Kraft ordered four Air Force aircraft to move into recovery positions in the Pacific for a possible splashdown some 800 km north-east of Hawaii. A naval destroyer and an oiler in the region were also ordered to stand by. Keenly aware of the situation, Cooper radioed that a decision needed to be made over whether to abort the mission or power down Gemini V’s systems and continue, to which Kraft told him to shut off as much as he could. All corrective instructions proved fruitless: neither the automatic or manual controls for the fuel cell’s oxygen tank heater would function. Nor could the heater itself, located in the adaptor section, be accessed by the crew. Cooper and Conrad even manoeuvred their spacecraft such that the Sun’s rays illuminated the adaptor, in the hope that it might stir the system back to life. It was all in vain.

By now, most of their on-board equipment – radar, radio, computer and even some of the environmental controls – had been shut down and, as Gemini V swept over the Atlantic on its third orbital pass, there was much speculation that a re-entry would have to be attempted before the end of the sixth circuit, since its flight track thereafter would take it away from the Pacific recovery area. Then, as the astronauts passed within range of the Tananarive tracking station in the Malagasy Republic, off the east coast of Africa, Cooper reported that pressures were holding at around 8.6 bars, suggesting, Kraft observed, that “the rate of decrease is decreasing”. As he spoke, the oxygen pressures dropped still lower, to just 6.5 bars, and fears were high that if they declined much further, Gemini V would need its backup batteries to support another one and a half orbits and provide power for re-entry and splashdown. The astronauts were asked to switch off one of the fuel cells to help the system and as they entered their sixth orbit the pressures levelled-out at 4.9 bars.

Capcom Jim McDivitt asked Cooper for his opinion on going through another day under the circumstances. “We might as well try it,’’ replied Cooper, but Kraft remained undecided. After weighing all available options, including the otherwise satisfactory performance of the cabin pressure, oxygen flow and suit temperatures, together with the prestige to be lost if the mission had to be aborted, he and his control team emerged satisfied that oxygen pressures had stabilised at 4.9 bars. If there were no more drops, Gemini V would be fine to remain in orbit for a ‘drifting flight’, staying aloft just long enough to reach the primary recovery zone in the Atlantic, sometime after its 18th orbit. Admittedly, with barely 11 amps of power, only a few of the mission’s 17 experiments could be performed, but Kraft felt ‘‘we were in reasonably good shape. . . we had the minimum we needed and there was a chance the problem might straighten itself out’’. As Cooper and Conrad hurtled over Hawaii on their fifth orbit, he issued a ‘go’ for the mission to proceed.

With the reduced power levels, the REP, which kept the spacecraft company up until its eighth orbit, was useless for any rendezvous activities. ‘‘That thing’s right with us,’’ Cooper told Mission Control during their sixth circuit of Earth. ‘‘It has been all along – right out in back of us.’’ Two orbits later, Conrad turned Gemini V a full 360 degrees, to find that the pod had re-entered the atmosphere to destruction. Nonetheless, Gemini V’s radar did successfully receive ranging data from the REP for some 43 minutes.

As the mission entered its second day, circumstances improved and oxygen pressures climbed. “The morning headline,” Kraft radioed the astronauts on 22 August, referring to a newspaper, “says your flight may splash down in the Pacific on the sixth orbit.” Having by now more than tripled that number of orbits, Conrad replied that he was “sorry” to disappoint the media. Despite the loss of the REP, on their third day aloft Cooper conducted four manoeuvres to close an imaginary ‘gap’ between his spacecraft and the orbit of a phantom Agena-D target. This ‘alternate’ rendezvous had been devised by the astronaut office’s incumbent expert, Buzz Aldrin. Cooper fired off a short burst from the aft-mounted OAMS thrusters to lower Gemini V’s apogee by about 22 km, then triggered a forward burn to raise its perigee by some 18 km and finally yawed the spacecraft to move it onto the same orbital plane as the imaginary target. One final manoeuvre to raise his apogee placed Gemini V in a co-elliptical orbit with the phantom Agena. Were it a ‘real’ target, he would then have been able to guide his spacecraft through a precise rendezvous. Such exercises would prove vital for Gemini VI, which was scheduled to hunt down a ‘real’ Agena-D in October 1965, and one of the greatest learning experiences, said Chris Kraft, ‘‘is being able to pick a point in space, seek it out and find it’’.

Notwithstanding the successes, the glitches continued. On 25 August, two of the eight small OAMS thrusters jammed, requiring Cooper to rely more heavily on their larger siblings and expend considerably more propellant than anticipated. It was at around this time that Gemini V broke Valeri Bykovsky’s five-day endurance record and Mission Control asked Cooper if he wanted to execute ‘‘a couple of rolls and a loop’’ to celebrate; the laconic command pilot, however, declined, saying he could not spare the fuel and, besides, ‘‘all we have been doing all day is rolling and rolling!’’ When the record of 119 hours and six minutes was hit, Kraft blurted out a single word: ‘‘Zap!’’ Gordo Cooper, with an additional 34 hours from Faith 7 under his belt, was now by far the world’s most flown spaceman. His response when told of the milestone, though, was hardly historic: ‘‘At last, huh?’’

The dramatic reduction of available propellant made the last few days little more than an endurance run. Kraft told the astronauts to limit their OAMS usage as much as possible and many of their remaining photographic targets – which required them to manoeuvre the spacecraft into optimum orientations – had to be curtailed. Still, a range of high-quality imagery was acquired. The hand-held 70 mm Hasselblad flew again to obtain photographs of selected land and near-shore areas and, of its 253 images, some two-thirds proved useful in post-mission terrain studies. These included panoramas of the south-western United States, the Bahamas, parts of south-western Africa, Tibet, India, China and Australia. Images of the Zagros Mountains revealed greater detail than was present in the official Geological Map of Iran. Cooper and Conrad also returned pictures of meteorological structures – including the eye of Hurricane Doreen, brewing to the east of Hawaii – together with atmospheric ‘airglow’. In addition, they took pictures of the Milky Way, the zodiacal light and selected star fields. Other targets included two precisely-timed Minuteman

missile launches and infrared imagery of volcanoes, land masses and rocket blasts.

The scientific nature of many of these experiments did not detract – particularly in the eyes of the Soviet media – from the presence of a number of military-sponsored investigations. Cooper and Conrad’s flight path carried them over North Vietnam 16 times, as well as 40 times over China and 11 times over Cuba, prompting the Soviet Defence Ministry’s Red Star newspaper to claim that they were undertaking a reconnaissance mission. The situation was not helped by President Johnson’s decision, whilst the crew was in orbit, to fund a major $1.5 billion Air Force space station effort, known as the Manned Orbiting Laboratory (MOL). Among the actual military experiments undertaken by Gemini V were observations of the Minuteman plume and irradiance studies of celestial and terrestrial backgrounds, together with tests of the astronauts’ visual acuity in space to follow up on reports that Cooper had made after his Faith 7 mission. Large rectangular gypsum marks had been laid in fields near Laredo, Texas, and Carnarvon, Australia, although weather conditions made only the former site visible.

Cardiovascular experiments performed during the mission would reveal that both men lost more calcium than the Gemini IV crew, although principal investigator Pauline Beery Mack expressed reluctance to predict a ‘trend’, since “a form of physiological adaptation may occur in longer spaceflight”. Medically, Chuck Berry’s main concerns were fatigue and his advice was that they get as much sleep as possible. ‘‘I try to,’’ yawned Conrad at one stage, ‘‘but you guys keep giving us something to do!’’ All in all, they managed between five and seven hours’ sleep at a time and expressed little dissatisfaction with Gemini V’s on-board fare: bite-sized, freeze-dried chunks of spaghetti and meatballs, chicken sandwiches and peanut cubes, rehydratable with a water pistol. An accident with a packet of shrimp, though, caused a minor problem when it filled the cabin with little pink blobs. Conrad even tried singing, out of key, to Jim McDivitt at one point.

Years later, Conrad would recall that the eight-day marathon was ‘‘the longest thing I ever had to do in my life’’. He and Cooper had spent the better part of six months training together, so ‘‘didn’t have any new sea stories to swap with one another… there wasn’t a whole lot of conversation going on up there’’. Nancy Conrad would recall her late husband describing how the confined cabin caused his knees to bother him – their sockets felt as if they had gone dry – and that he would have gone ‘‘bananas’’ if asked to stay aloft any longer. (Ironically, on two future missions, Conrad would stay aloft for much, much longer. . . but on those occasions, his tasks would include a couple of meandering trots around the lunar surface and floating inside a voluminous space station.) He found it hard to sleep, hard to get comfortable and the failures meant he and Cooper spent long periods simply floating with nothing to do. After the flight, he told Tom Stafford that he wished he had taken a book, and this gem of experience would be noted and taken by the crew assigned to fly the 14-day mission.

Nancy Conrad described Cooper’s irritation at losing so much of his mission. He was far from thrilled that the two main tasks for Gemini V, rendezvous and long – duration flight, were becoming little more than ‘‘learning-curve opportunities’’ and suggested throwing an on-board telescope in the Cape Kennedy dumpster when it twice refused to work. Later, when the spacecraft was on minimum power and the astronauts were still expected to keep up with a full schedule, Cooper snapped “You guys oughtta take a second look at that!” As for physical activity, he grimaced that his only exercise was chewing gum and wiping his face with a cleansing towel.

On the ground, Deke Slayton was concerned that such an attitude would not help Cooper’s reputation with NASA brass. Indeed, Gemini V would be his final spaceflight and, although he would later complain bitterly about ‘losing’ the chance to command an Apollo mission, some within the astronaut corps would feel that Cooper’s performance and strap-it-on-and-go outlook had harmed his career. Tom Stafford was one of them. ‘‘Gordo… had a fairly casual attitude towards training,’’ he wrote, ‘‘operating on the assumption that he could show up, kick the tyres and go, the way he did with aircraft and fast cars.’’

To spice matters up still further, worries about the fuel cells continued to plague Gemini V’s final days. Their process of generating electricity by mixing hydrogen and oxygen was producing 20 per cent too much water, Kraft told Conrad, and there were fears that the spacecraft was running out of storage space. This water excess might back up into the cells and knock them out entirely. In order to create as little additional water as possible, the astronauts powered down the capsule from 44 to just 15 amps and on 26 August Kraft even considered bringing them home 24 hours early, on their 107th orbit. However, by the following day, the water problem abated, largely due to the crew drinking more than their usual quota and urinating it into space, and a full-length mission seemed assured.

Eitherway, they had long since surpassed Bykovsky’s Vostok 5 record. In fact, by the time Cooper and Conrad splashed down, they would have exceeded the Soviets on several fronts: nine manned missions to the Reds’ eight, a total of 642 man-hours in space to their 507 and some 120 orbits on a single mission to their 81. At last, after eight years in the shadows – first Sputnik, then Gagarin, Tereshkova, Voskhod 1 and Leonov – the United States was pulling ahead into the fast lane of the space race. When it seemed that Gemini V might come home a day early and miss the scheduled Sunday 29 August return date, mission controllers in Houston even played the song ‘Never on Sunday’, together with some Dixieland jazz.

The astronauts also had the opportunity on the last day of the mission to talk to an ‘aquanaut’, Aurora 7 veteran Scott Carpenter, who was on detached duty to the Navy. Carpenter, who had broken his arm in a motorcycle accident a year before and been medically grounded by NASA, was partway through a 45-day expedition in command of Sealab II, an underwater laboratory on the ocean floor, just off the coast of La Jolla, California. The Sealab effort, conceived jointly by the Navy and the University of California’s Scripps Institution of Oceanography, sought to discover the capacity of men to live and work effectively at depth. In doing so, Carpenter became the first person to place ‘astronaut’ and ‘aquanaut’ on his career resume. Yet, unlike Cooper and Conrad, his chances of returning to space were non­existent. He had not impressed senior NASA managers with Aurora 7 and, indeed, the partial success of an operation to repair the injury to his arm meant he would remain grounded anyway. He resigned from NASA in early 1967.

The music, the chat with Carpenter and even Conrad’s dubious singing did little

to detract from the uncomfortable conditions aboard the capsule. As they drifted, even with coolant pipes in their suits turned off, the two men grew cold and began shivering. Stars drifting past the windows proved so disorientating that they put covers up. Sleep was difficult. Chuck Berry had wired Conrad with a pneumatic belt, a blood-pressure-like cuff, around each thigh, which automatically inflated for two minutes of every six throughout the entire mission. The idea was that, by impeding blood flow, it forced the heart to pump harder and gain its much-needed exercise. Berry felt that if Conrad came through Gemini V in better physical shape than Cooper, who did not wear the belt, a solution may have been found for ‘orthostatic hypotension’, the feelings of lightheadedness and fainting felt by some astronauts after splashdown.

For the two astronauts, that splashdown could not come soon enough. By landing day, 29 August, their capsule had become cluttered with rubbish, including the litter of freeze-dried shrimp, which had escaped earlier in the mission. The appearance of Hurricane Betsy over the prime recovery zone prompted the Weather Bureau to recommend bringing Gemini V down early and Flight Director Gene Kranz agreed to direct the Lake Champlain to a new recovery spot. At 7:27:43 am, Cooper fired the first, second, third, then fourth OAMS retrorockets, then gazed out of his window. It felt, he said later, as if he and Conrad were sitting ‘‘in the middle of a fire’’. Since it was orbital nighttime, they had no horizon and were entirely reliant upon the cabin instruments to control re-entry. In fact, Gemini V remained under instrument control until they passed into morning over Mississippi.

Cooper held the spacecraft at full lift until it reached an altitude of 120 km, then tilted it into a bank of 53 degrees; whereupon, realising that they were too high and might overshoot the splashdown point, he slewed 90 degrees to the left to create more drag and trim the error. Although experiencing a dynamic load of 7.5 G after eight days of weightlessness, the astronauts did not, as some had feared, black out. The parachute descent was smooth. No oscillations were evident and the 7:55:13 am splashdown, though 170 km short of the planning spot, was soft. As would later be determined, the computer had been incorrect in indicating that they would overshoot. A missing decimal point in a piece of uplinked data had omitted to allow for Earth’s rotation in the time between retrofire and splashdown. In fact, Cooper’s efforts to correct the false overshoot had progressively drawn them short of the recovery zone. ‘‘It’s only our second try at controlling re-entry,’’ admitted planning and analysis officer Howard Tindall. ‘‘We’ll prove yet that it can be done.’’

Gemini V had lasted seven days, 22 hours, 55 minutes and 14 seconds from its Pad 19 launch to hitting the waves of the western Atlantic and the crew was safely aboard the Lake Champlain by 9:30 am. With the exception of the failed REP rendezvous, and one experiment meant to photograph the target, all of Cooper and Conrad’s objectives had been successfully met. Yet more success came when Chuck Berry realised that, despite the days of inactivity with little exercise aboard the capsule, the astronauts were physiologically ‘back to normal’ by the end of August, clearing the way for Frank Borman and Jim Lovell to attempt a 14-day endurance run on Gemini VII in early 1966. First, though, Wally Schirra and Tom Stafford would fly Gemini VI for one or two days in October and complete the first rendezvous with an

Agena-D target. The mission – or, rather, missions – that would follow would snatch victory from the jaws of defeat and set aside another obstacle on the path to the Moon. But not before suffering a major setback of its own.