Category BOLD THEY RISE

Science on the Shuttle

As the nation’s Space Transportation System, the primary goal of the Space Shuttle was just that—to transport people and cargo back and forth be­tween the surface of Earth and orbit around the planet. However, it was quickly obvious that the shuttle had even greater potential.

In 1973 and ’74, the Skylab space station had demonstrated the value of conducting scientific research in Earth orbit. From the beginnings of the Space Shuttle program, plans had been that the vehicle would support a space station that would continue, and build further upon, the work done on Skylab. However, the realities of budgetary constraints meant that the two could not be developed at the same time and that the space station would have to wait.

In the meantime, though, the Space Shuttle itself would provide a stop­gap measure— a pressurized module placed inside the orbiter’s cargo bay could be roomy enough to serve as an effective orbital science laboratory.

STS-9

Crew: Commander John Young, Pilot Brewster Shaw, Mission Special­ists Owen Garriott and Bob Parker, Payload Specialists Ulf Merbold (Germany) and Byron K. Lichtenberg

Orbiter: Columbia

Launched: 28 November 1983

Landed: 8 December 1983

Mission: First flight of the Spacelab laboratory module

Launched in November 1983, STS-9 marked a new step forward for the Space Shuttle program, while also hearkening back to the first four mis­sions. The first mission to fly the Spacelab science module, STS-9 was both a demonstration and an operational flight. The primary purpose was to make sure that the Spacelab module worked properly, but that was achieved

by conducting a full complement of scientific experiments. According to STS-9 mission specialist Owen Garriott, who was a member of nasa’s first group of scientist-astronauts and of the second crew of Skylab, the mission of Spacelab I included experiments in biomedicine, astronomy, fluid phys­ics, materials processing, and atmospheric sciences.

A rack of fluid physics equipment enabled the crew to inject liquids, shake them and rotate them, combine them and stretch them, in order to study the small forces associated with the surface tension and the even smaller forces that are associated with fluids that cannot be as easily demonstrated or measured in a one-g environment. An adjacent double rack was dedi­cated to materials science. The racks included furnaces in which material samples could be heated and resolidified. Other equipment focused light on a crystal, melted it, and then allowed it to solidify more carefully. Spec­trometers flown on the mission allowed the crew to conduct stellar astron­omy and air-glow research through the orbiter’s windows.

The variety of the experiments was what drew Garriott to the flight. Gar­riott, who spent more than fifty-nine days in space as a science pilot astro­naut aboard Skylab, explained that for a mission like Skylab or STS-9 that is focused on interdisciplinary scientific research, the ideal crew member is a scientific generalist, with interest and ability in multiple areas, rather than someone who specializes in one field. “My interest is interdisciplinary work, so I found [the mission] quite interesting, and I think it does relate to the fact that you need an interdisciplinary background to try to conduct experiments for all of these pis [principal investigators]. You obviously can’t have a representative for each pi there, and you need somebody who has got some degree of competency in all of the variety of areas.”

Before the flight assignment, during the period between Skylab and shut­tle, Garriott had gone into management, becoming director of science and applications at Johnson Space Center. He was still a member of the astro­naut corps but was no longer actively involved with the Astronaut Office.

I had my hands full two buildings away. I think I still talked to the folks in the Astronaut Office all the time. I was still flying airplanes. I might have been going over to those weekly meetings as well. But I did not have any act­ing role in the Astronaut Office from the standpoint of getting ready for shut­tle or anything else. I was expecting to stay in the Science and Applications

Science on the Shuttle

26. In the Spacelab on the first Spacelab mission are Robert Parker, Byron Lichtenberg, Owen Garriott, and Ulf Merbold. Courtesy nasa.

Directorate until another flight opportunity came along. I’d always intended to return to the Astronaut Office full time as soon as there was the first flight opportunity available.

Garriott started training for the shuttle in 1978, when nasa believed the mission was still only about three years away. “We had a lot of training to do,” Garriott recalled, “because there were something like sixty experiments on board, half from Europe, half from the U. S., roughly. And we visited almost every pi at their facilities and talked with them about how to oper­ate it and the hardware development.”

During that period the crew was very involved in the development of the Spacelab hardware, particularly the controls and displays the astronauts would use to interface with the equipment. “If you don’t understand how [something] works, if it’s not something that’s human friendly, you can waste a lot of time or make a lot of mistakes,” Garriott said. “That was the one thing that supposedly the crew members were more expert in, the in­terface between man and machine.”

Delays meant that Garriott did not make his second spaceflight until a full decade after his first, but he said it was worth it. “That ten-year delay is longer than I expected, but it’s a decision that I really made when I went

over to Science and Applications in the first place. I wanted to come back and fly again. And so I was just anxious to get back and fly again.”

The variety of the experiments meant that the crew members had to spend a substantial amount of time learning about the experiments and working on the ground with the principal investigators with whom they would be interacting from orbit. There had been fewer disciplines represent­ed on Skylab, Garriott recalled, so crew members did not have to travel as much or as far to talk with the principal investigators. “In fact, for most of the time, they came to jsc for our solar physics training and for our bio­medical training.” On Spacelab, the investigators were all over the world, with experiments from around the United States, from countries in Europe, and one from Japan. “We were very international and traveled all over the world in order to talk with the investigators about the conduct of their ex­periments. I feel so fortunate to have had the chance to work in all of these different disciplines with the real world’s experts and to learn from them.”

Garriott also felt fortunate to work with the other astronauts on STS-9, including Commander John Young.

I’ve always gotten along extremely well with my fellow crewmates. John Young, who was the commander of Spacelab I, is not an Alan Bean [commander ofGar – riott’s Skylab crew]. He’s motivated differently, different personality, [a] standard sort ofaprototypical test pilot. I was a little concerned about that, since we were bringing in the first foreign international. How well is that going to work? It turns out it worked extremely well, for which I give John Young a lot ofcredit. I really enjoyed having John as the commander ofour flight. And I think after we came back he had no better friend on board our flight than Ulf Merbold. They still often talk. Whenever Ulf comes back, I think they get together for dinner.

While the commander’s primary responsibility in orbit was the operation of the vehicle and oversight of the crew, Garriott noted that Young made several other contributions to support the scientific portions of the mis­sion. For example, when the mounting apparatus for a camera used for a vestibular experiment broke, rather than allowing thousands of film frames to go to waste, Young took the camera and film to the flight deck and spent time taking Earth-observation photographs.

“John really jumped in and assisted in the conduct of the science,” Gar­riott commented. “Before flight there was some little concern about that,

that whether or not a standard prototypical test pilot would enjoy and par­ticipate. But we all really enjoyed having him on board. He did his share, as did the rest of us.”

sts-9 was the first nasa flight to include payload specialists, astronauts who were not career nasa employees. Ulf Merbold, the first internation­al astronaut to fly on the shuttle, was a West German astronaut from the European Space Agency, and Byron Lichtenberg was a researcher from the Massachusetts Institute ofTechnology (mit).

Garriott described the feelings of the career astronauts toward the payload specialists as “uncertainty” and compared it to when his class of scientist – astronauts joined the all-pilot astronaut corps.

When we came into the Astronaut Office, I can imagine that the pilots were thinking, “Geez, these old fuzz-hair university types, can they hold their own here, do they know what’s going on?" In the same way, these [payload spe­cialists] are really university types here. Are they really motivated to fly, can they hold their own, and so forth? And I very quickly found out that yes, in­deed, they could. They were on par with all of us. We were very much of the same kind of breed, in my opinion. We got along extremely well. Everybody got along fine. And after thirty years, they’re still some of my best friends. So it was a very pleasant and positive experience for both ms and ps [mission spe­cialists and payload specialists] as far as Spacelab I is concerned. I think that is generally true for most of the flights, though not necessarily in each individ­ual case. When you get up to twenty, thirty people, you’re bound to find some rough corners somewhere. I can say on Spacelab I, it was a remarkably posi­tive experience for all of us.

While Garriott did not recall any resentment over payload specialists tak­ing slots that otherwise might have gone to career astronauts, he also not­ed that in the case of STS-9, the payload specialists had trained for as long as the newest nasa astronauts had been in the corps.

Brewster Shaw, the pilot for STS-9, said that, being in a crew that was mostly rookie space flyers, it was interesting working with veteran astro­nauts Garriott and Young.

John and Owen were the only two guys who had flown. [Robert] Parker had never flown. I’d never flown, and Ulf and Byron had never flown before. So John and

Owen were the experienced guys, and they kind of were the mentors of the rest of us. It was fun to watch Owen Garriott back in the module, because you could tell right from the beginning he’d been in space before, because he knew exactly how to handle himself how to keep himselfstill, how to move without banging all around the other place. And the rest of us, besides he and John, the rest of us were bouncing off the walls until we figured out how to operate. But Owen, it was just like, man, he was here yesterday, you know, and it really had been years and years.

Garriott was among the very small number of astronauts to have the experience of flying on both the Saturn IB rocket and the Space Shuttle. “The launch phase is remarkably similar, because both vehicles have a thrust

which is just a little bit more than the weight of the vehicle. You start out

at very low acceleration levels,” Garriott explained, comparing the g-forces of the vehicles’ initial accelerations to what one would experience in a car.

It keeps building up a little bit, but the important thing is it’s steady. When you accelerate in a car, you accelerate for three or four seconds and then you reach the speed you want to be at. Here, you keep on accelerating. So you keep up an acceleration which is increasing up to maybe three or four gs. But it takes about eight and a half minutes, nine minutes, to get to space on either one of the two vehicles. As you lift off there’s quite a bit of vibration with either vehicle. Once you get above about one hundred thousand feet, first ofall, you stage, you switch to a smaller engine on a second stage, and you’re above most of the atmosphere so the vibration diminishes. Then it just becomes a nice steady push, and you con­tinue that push at about three to four gs, all the rest of the way till you’re into orbit. All of a sudden you’re bouncing up against your straps. So the launch is remarkably the same, in my view, between a Saturn and a shuttle.

With the development, training, and launch complete, it was time for the real business of STS-9 to begin. The crew worked for twenty-four hours a day, divided into two shifts. The commander and pilot each were assigned one shift, and the scientists were split between them. Each crew member was on duty for about twelve hours, with a brief handoff between shifts. The astronauts then had twelve hours off duty, during which to sleep, eat, get ready, and enjoy the experience of being in space.

Shaw remembered that he and Young would spend the majority of their work shifts on the orbiter’s flight deck.

You didn’t want to leave the vehicle unattended very much, because this is still STS-9, fairly early in the program. We hadn’t worked out all the bugs and ev­erything, and neither John nor I felt too comfortable leaving the flight deck un­attended, so we spent most of our time there. We had a few maneuvers to do once in a while with the vehicle, and then the rest of the time you were moni­toring systems. After a few days of that, boy, it got pretty boring, quite frankly. You spent a lot of time looking out the window and taking pictures and all that. But there was nobody to talk to, because the other guys were back in the back end in the Spacelab working away, and, you know, you just had this, “Gosh, I wish I had something to do, " kind of feeling.

While he and Young were not heavily involved in the research part of the mission, Shaw did recall being part of one experiment, designed to study how humans adapted to working in space.

I did “Helen’s balls.". . . Helen was a principal investigator, and she had a bunch of little yellow balls that were different mass. . . . Since there’s no weight, there’s only mass, in zero g we had to try and differentiate between the mass of these balls. You would take a ball in your hand and you would shake it and you would feel the mass of it by the inertia and the momentum of the ball as you would start and stop the motion. Then you’d take another one and you’d try and differentiate between [them], and eventually you’d try and rank [the] order of the balls. . . . And, quite frankly, that’s the only experiment I remember doing.

Garriott used his second flight as an opportunity to make history in a way that had a lasting legacy for the space program. When Garriott was younger, his father took an amateur radio class with him, and the two became ham operators. After he was assigned to sts-9, Garriott got permission to take an amateur radio rig with him on Columbia and was able to make contact with people on the ground. “I’d say fifteen to twenty hours was reserved for use of the ham equipment. . . after the working day was over,” Garriott re­called. “I found that quite interesting and enjoyable, and it turned out to be a very positive thing from the standpoint of the ham radio community.”

While Garriott mostly talked with whomever he happened to get in con­tact with, he had prearranged a few conversations. “There were a few high – profile people that that’s the only way you’ll be able to get in touch with them. Normally, when you just want to talk to anybody, you use a call signal that’s called ‘cq.’ And if you call CQ from space, you’ll probably get a hun­dred answers, so you’re way overloaded. So if you want to [reach] one par­ticular person, you have to specify a time and frequency that’s kept private.” One of the preplanned calls was to Garriott’s hometown of Enid, Okla­homa. “I made arrangements to talk with them and my mother. . . with one of the local hams on the radio,” he said. “Another one was [Senator] Barry Goldwater, who is a very well-known ham in the Senate and always looked after amateur radio very well from that position. And another inter­nationally known, great enthusiast was King Hussein from Jordan.”

For the hundreds of other people Garriott talked to in addition to those preplanned calls, he used a tape recorder to log the contacts. “Everyone I could discern answering my call, I would repeat back their call sign, as many as possible. And then I listened to every one of those tapes. And I had an­other friend go through the whole list, and we responded to everyone whose call sign we could pick out of there. So there were, I don’t know, four or five hundred people that all received a card called a QSL, that’s another special symbol that’s used by hams as a symbol of a contact between two people.” Garriott’s initiative to use amateur radio on STS-9 created a lasting leg­acy. Almost thirty additional shuttle flights carried ham radio and used it to talk to schoolchildren and other groups, allowing students to ask ques­tions of astronauts in orbit during organized events. Years later, amateur ra­dio became a fixture on the International Space Station, with its first crew conducting a ham radio conversation with a school within weeks of board­ing the station. Thousands of students have had the chance to talk to orbit­ing astronauts as a result of the amateur radio contacts begun by Garriott.

That legacy paid off for Garriott himself in a very special way many years later. In 2008 Garriott’s son, Richard, became the first second-generation American space flyer as a spaceflight participant on a Russian Soyuz flight to the International Space Station. While there, Richard was able to use the equipment on the Space Station to talk to his father on the ground. As unique as that experience was, it had an even deeper personal significance for the elder Garriott. “My father and I got our ham licenses together back in 1945, and his call sign was W5KWQ. Mine was W5LNL. Now, normally, when a person dies, they reserve that for a little while, but then they’ll even­tually reissue the call sign. So after my father died, which would have been back in ’81, it hadn’t been reissued yet. So my son got his license [in 2006], and they allowed him to have his grandfather, my father’s, call sign again.” During his son’s flight, Garriott was able to respond to a contact from his father’s call sign, now used by his son.

Garriott recalled another amusing communications-related incident dur­ing the STS-9 flight. Given the historic nature of the first flight of an in­ternational astronaut, a video link was set up between the shuttle, Presi­dent Ronald Reagan, and West German chancellor Helmut Kohl. The link was to include Commander (and moonwalker) Young, German astronaut Merbold, and the U. S. payload specialist, Lichtenberg. In order to save the time of the three crew members to be featured, the two mission special­ists and the pilot were asked to help set up the link and then turn it over to the others. In response to being treated as second-class astronauts, the trio made a silent protest—when the link was established for testing, the video showed Garriott, Parker, and Shaw in a classic “see no evil, hear no evil, speak no evil” pose, with one each covering his eyes, ears, and mouth. “What we were told was that was a hilarious, humorous scene in the con­trol room and that it got into Time magazine.”

While Garriott’s launch experiences were fairly similar between the Saturn IB and the Space Shuttle, his landing experiences were not. Initially, he ex­plained, the return to Earth begins similarly on both vehicles. A slight slow­ing causes the vehicle to dip into the atmosphere, creating an unforgettable glow as the air around the spacecraft heats up. “On the shuttle I happened to be sitting right beside the side window on the mid-deck, so you see all of the blue and yellow flames coming by. Then you get a little lower, it will turn to orange and reddish flames as the temperatures drop down a little bit.”

At that point, the differences between the two reentries begin to be obvious.

As you really get lower, on the command module you have to pop drogue chutes to orient your spacecraft. When you get down to about ten thousand feet, I believe it is, the main chutes come out, and then you finally get down to a big splash in the water. That’s quite different than coming down in a glider. You’ve been used to seeing the rate at which the Earth moves by [from space]. Once you get down to a lower altitude, the Earth starts moving by faster and faster, because you’re lower, you’re closer to it. So it’s interesting to see those two comparisons.

The landing of STS-9 was more than a little unusual. According to Shaw, “We had another lesson on the landing of STS-9, and the lesson was, never let them change the software in the flight control system without having adequate opportunity to train with it.”

The shuttle’s flight control software interprets what actions it needs to take based on the inputs the commander makes on the stick, and how it responds depends on what point it is at in the mission. The settings that Young and Shaw used during training were replaced for the actual flight, which changed how the vehicle handled.

“I don’t remember if we knew about that or didn’t know about that,” Shaw said,

but certainly when John started to de-rotate the vehicle, it responded different­ly than he had trained on. So here we are. John’s flying the vehicle. I’m giving him all the altitude and airspeed calls and everything, and you feel this nice main gear gently settling onto the lake bed. . . . There were only two of us on the flight deck, as I recall, because we still had the ejection seats in Columbia at that time. They hadn’t been taken out, so there was no room for another seat. So. .. the other four were on the mid-deck, and you hear this, “Yay!" and clap­ping when the main gear touched the ground very gently. Then John gets the thing de-rotated and we’re down to about 150 knots or so when the nose gear hits the ground, and it goes “smash!" So it changes from this “Yay! Yay!" to “Je­sus Christ! What was that?" That was just really funny, and I got all of that on tape, because I had a tape recorder going. And poor John, he was embarrassed because of this, the way the nose gear hit down, but it wasn’t his fault. They had changed this thing without him being able to practice using the new flight control system. So that was a good lesson.

The crew learned the next day that Columbia had also suffered yet an­other problem during the reentry and landing—a fire in the vehicle’s aux­iliary power units. “We had one apu shut down, and then when we shut the other two apus down, normally after landing, it turns out one of them was also on fire,” Shaw explained.

The reason the first one shut down was it was on fire. . . and it automatically shut itself down. . . . The next one didn’t shut down until we actually shut it down. But there were two of them that were burning. . . . So we had a fire out­side the APUs that when we shut them down and shut the ammonia off to them, the fires went out. So we had some damage back there, but the fires stopped.

But we didn’t know anything about that till the next day. I got this call and John says, “Hey, did you know that the apus were burning?" "No, I sure didn’t. ”

The flight also had two of the general-purpose computers (gpcs) fail. “That was an interesting thing, too,” Shaw recalled.

John and I were in the de-orbit prep, . . . and we were reconfiguring the GPCs and the flight control systems and the rcs [reaction control system] jets and stuff. . . . About the time that we were reconfiguring the computers, we had a couple of thruster ^firings, and the big jets in the front fired and it’s like these big cannons—boom! boom!—and it shocks the vehicle. You know, you really can feel it if you’re touching the structure. So we had one of these firings and we got the big “Xpole fail” on the CRT, meaning the computer had failed. This is the first computer failure we had on the program. Our eyes got about that big. So I get out the emergency procedure checklist and. . . we started going through the steps and everything. And in just a couple of minutes we had another one fail the same way. . . . So now we were really interested in what was going on. We ended up waiving off our de-orbit at that time. . . . The ground decided, no, we’re going to wait and try to figure out what’s going on with these computers.

When Young’s shift ended, with nothing more he could do to address the problem and wanting to be rested for the landing, he went down to the mid-deck to take a nap, leaving Shaw “babysitting” on the flight deck.

Well, during that timeframe, all of a sudden there starts this noise, bang, bang, bang, bang, bang, bang, bang, bang. The next thing, one ofour three imus [in­ertial measurement units] . . . failed and we couldn’t recover it. It turned out its gimbal was failing and it was beating itself to death against the gimbal stop, and that was the banging noise. After a few hours, John comes back upstairs and says, “You know, I really appreciate you guys making all that banging noise when I’m trying to sleep down there. ” I said, “Jeez, John, I’ve got some bad news. Man, we lost an imu. ” And John’s eyes get this big again, because we’ve had two gpc failures and now an imu failure. Anyhow, we got through all that and we entered and landed, and when the nose gear slapped down, one of the gpcs that had recovered failed again. One of them didn’t recover and we flew down with one less computer, but that computer failed again, and that’s why I reconfig­ured flight control systems, as I remember now, because of that computer failure.

Despite the drama of the landing, once Columbia was successfully and safely on the ground, Shaw and the agency were pleased with the results of the shuttle’s first dedicated scientific flight. “We learned a lot from that flight, a tremendous amount. Seventy-seven different investigations, as I recall, on that mission. It was a tremendous success.”

STS-51B

Crew: Commander Bob Overmyer, Pilot Fred Gregory, Mission Special­ists Don Lind, Norm Thagard, and Bill Thornton, Payload Specialists Lodewijk van den Berg and Taylor Wang Orbiter: Challenger Launched: 29 April 1985 Landed: 6 May 1985 Mission: Second Spacelab flight

The next Spacelab flight, 51B, came in April 1985. And there may have been no one more excited about it than astronaut Don Lind, who had been selected in NASA’s fifth group of astronauts in 1966. “I set a record. No one has waited for a spaceflight longer than I have,” Lind said. “I hope nobody ever has to do that. But with the six and a half years I spent in training for two flights that didn’t fly, and then the delays in getting the shuttle pro­gram going, and with the [Apollo i fire, there were long delays, and so it was nineteen years before I got to fly.”

For Fred Gregory, the pilot of the mission and one of the astronauts se­lected in the class of ’78, the wait wasn’t nearly so long, but it was none­theless a huge honor to finally be selected for a crew. “There was shifting of launch times, because we were on Spacelab III and I know we launched before Spacelab II and after Spacelab I. I think they had some payloads that they wanted to deploy quickly, and so the laboratory missions were kind of put in a kind of a second category for priority. But, you know, the date wasn’t important at the time.”

Gregory found it interesting that the overwhelming majority of his train­ing was for a very small part of the mission. Over three-fourths of his train­ing focused on the eight and a half minutes of launch and hour of entry, out of a week-long mission. While the proportions may have seemed odd, it was necessary to make sure the flight deck crew was ready to perfectly ex­ecute the tasks required of it no matter what happened.

It’s like a ballet, you know, without music: individual but coordinated activities that resulted in the successful accomplishments ofeach of these phases, regardless of the type offailure or series offailures that this training team would impose on you. So that’s what we trained for. There were two thousand or so switches and gauges and circuit breakers, any number of which we would involve ourselves with during these two phases, ascent and entry. So the intent was for us to learn this so well, understand the system so well, that we could brush through a failure scenario and safe the orbiter in the ascent such that we could get on orbit and then have time to discuss what the real problem was and then allow you to correct it.

Eventually, it was time to move from the simulations to the real thing. As with sts-9, two shifts worked around the clock. “While each shift worked, the other shift slept,” Gregory said.

We had enclosed bunks on the mid-deck of the orbiter, and that’s where the off shift would sleep, so we never saw them really…. Each shift had its own area of interest, so there was really no competition between the two of them. . . . There were really about four hours a day there was an interaction between the two. It would be the two-hour postsleep transition, when one is just waking up. A shift is waking up, and they are picking up the ball, so to speak, from the other shift. The other shift then prepares to go to bed.

As pilot, Gregory’s responsibilities in orbit mainly focused on the support systems that kept the orbiter functioning. He was technically in charge of one shift, but since the scientists were largely capable of carrying out their duties on their own, his support of them centered around maintaining the vehicle. “During that mission there were very few problems, and those that we had were very minor. So not only did I monitor the shuttle, but I also had a great deal of time to learn about the Earth, and then spent a lot of time looking into deep space. We were in a high-inclination orbit, 57D de­grees, and so it gave us an excellent view of a great part of the Earth.”

The crew was joined by some rather unusual travel mates, recalled Don Lind:

We had the first laboratory animals in space, and Bill Thornton had to wor­ry about them on one shift, and Norm would worry about them on the other shift. . . . We had two cute little squirrel monkeys and twenty-four less-than – cute laboratory rats.

Science on the Shuttle

27. The Spacelab in the cargo bay of the Space Shuttle. Courtesy nasa.

The squirrel monkeys adapted very quickly. They had been on centrifuges. They had been on vibration tables. So they knew what the roar and the feeling of space was going to be like. Squirrel monkeys have a very long tail, and ifthey get excited, they wrap the tail around themselves and hang on to the tip of the tail. Ifthey get really excited, they chew on the end of their own tail. By the time we.. . activat­ed the laboratory, which was about three hours after liftoff they were now adjust­ed. They had, during liftoff apparently chewed off about a quarter of an inch of the end of their tails, but they were adjusted and just having a ball. I kept saying, “Let’s let one of them out." “No, no, cant do that. We’d never catch him again."

While the monkeys adapted quickly, the laboratory rats were somewhat slower. “The laboratory rats were not quite as savvy as the monkeys,” Lind said. “They had also been on vibration tables and acoustical chambers and that sort of thing, but they hadn’t learned that this was going to last awhile, and when

we got into the laboratory, they were hanging on to the edge of the cage and looking very apprehensive. After about the second day, they finally found out if they’d let go of the screen, they wouldn’t fall, and they probably enjoyed the rest of the mission. But they were slower in adapting. No big problem.”

Gregory noted that he and Overmyer had the “privilege” of helping with the monkeys at one point, doing cleanup work capturing debris from the food and waste escape in the holding facility.

These rhesus monkeys that we had were extremely spoiled. I think that the en­vironment that they had come from before they came on the orbiter was a place where they received a lot ofattention from the caregivers there. Norm and I would look back into the Spacelab, and we would see Bill Thornton attempting to get these monkeys to do things, like touch the little trigger that would release the food pellets. And I could tell. .. watching them that they expected Bill to do that for them. So we looked back there one time, and we could see that kind of the roles were reversed, that Bill was actually doing antics on the outside of the cage and the monkeys were watching. We almost joked sometimes that they started laugh­ing, and they went back and ate. It was an interesting dynamic to watch Bill Thornton wrestle with or react with the monkeys. It was quite an act back there.

Lind recalled that the substantial amount of automation involved in the Spacelab experiments provided him with a unique opportunity to do an ex­periment of his own. When he realized that a good bit of his duties would consist of regularly checking on automated equipment, he and a partner proposed an experiment to look at the aurora from space. Prior to his mis­sion, the aurora had been photographed from space by slow-scan photom­eters, which give a blurred picture. On Skylab, Owen Garriott had taken a limited number of photographs of the aurora on the horizon.

“There were a few pictures, but not many,” Lind explained.

So Tom and I started thinking about how we could do this. The first thing we wanted were high-time-resolution pictures of the aurora made with a TV cam­era. So we started looking around. What TV system could we get that would be sensitive enough in such a low light level? It turned out that the tv camera that was already on the Space Shuttle was as good as any TV camera we could have bought in the world. But we had to take off the color wheel and photograph in black and white instead of color. So we asked and got one of the TVs modi­fied. Because [we would be doing that only] in black and white, you want to take some still photographs in color to document what color the auroral light is, since that will identify what particles are emitting the light. So we started to look around for an appropriate camera and camera lens. It turned out that the camera that we already had on board and the lens we already had on board were, again, as good as we could have gotten anywhere. NASA only had to buy three rolls of film, special, sensitive color film. So this experiment cost NASA thirty-six dollars, and it’s the cheapest experiment that has ever gone into space. It was very satisfying to us that we made some discoveries on thirty-six dollars. We claimed that we could do more science per dollar per pound than anybody else in the space program.

The experiment proved to be not only cheap, but also effective, revealing a new component of aurora formation.

Despite the busy shift schedule and the additional experiment Lind had proposed to fill the extra time, he eventually found some time to appreci­ate the view from the orbiter’s flight deck windows.

It was like a Cinerama presentation. Both my wife and I are amateur oil paint­ers. … I thought, “Could I ever paint that?"The answer is absolutely not. Grum – bacher [art materials company] doesn’t make a blue that’s deep enough for the great ocean trenches. You look out tangentially through the Earth’s horizon, and you see—I was quite surprised—many different layers of intense blue colors, about twenty to twenty-two different layers of cobalt and cerulean and ultra­marine and other shades, and then the blackest, blackest space you can imag­ine. When you go over the archipelagoes and the atolls and islands in the Pa­cific and down in the Bermudas, you see the water coming up toward the shore from the deep trenches, and it appears as hundreds of shades of blue and blue green up to a little white line, which is the surf, and another little brown line, which is the beach. Nobody will ever paint that. It’s magnificent.

Lind was so overwhelmed by the beauty of the view that it brought tears to his eyes, which he discovered was also a very different experience in orbit.

In space, tears don’t trickle down your cheeks; that’s caused by gravity. In space the tears stay in the eye socket and get deeper and deeper, and after a minute or two I was looking through a half inch of salt water. I thought, “Ooh, this is like a guppy trying to see out of the top of the aquarium." Simultaneously with this, there’s this sense of incredible beauty. But then I had a spiritual feeling, be­cause several scriptures popped into my mind. You know, the nineteenth Psalm, “The heavens declare the glory of God." One of the unique Mormon scriptures is, “If you’ve seen the corner of heaven, you’ve seen God moving in his majesty and power." In the book of Romans, it says that the righteous will be coinher­itors with Jesus Christ of all this. I thought, you know, “This must be the way the Lord looks down at the Earth." Because from space, you cant see any gar­bage along the highway. You cant hear any family fights. You see just the beau­ty of how the Lord created this Earth, and that was a very spiritual, moving experience, along with the aesthetics, the physical beauty. I’ll always remem­ber that special feeling, besides the technical satisfaction of a successful mission.

STS-51F

Crew: Commander Gordon Fullerton, Pilot Roy Bridges, Mission Spe­cialists Story Musgrave, Tony England, and Karl Henize, Payload Spe­cialists Loren Acton and John-David Bartoe

Orbiter: Challenger

Launched: 29 July 1985

Landed: 6 August 1985

Mission: Third flight of the Spacelab laboratory module

The third Spacelab flight, 51F, was commanded by Gordon Fullerton, marking his return to space after serving as pilot of the final demonstration flight, sts-4. “It was a great mission,” Fullerton commented.

It really was. Some of the missions were just going up and punching out a sat­ellite, and then they had three days with nothing to do and came back. We had a payload bay absolutely stuffed with telescopes and instruments. We had the instrument-pointing system that had never been flown. We had the idea of let­ting a satellite go and then flying this precise orbit around it and then going back and getting it. So, all kinds of new things, which took a lot of work to write the checklists for, write the flight plan, and so we spent a year and a halfdoing that.

After all the preparation, the mission had a rather inauspicious begin­ning. All three Space Shuttle main engines lit properly on schedule seconds before the scheduled launch, but then one failed before the solid rocket mo­tors were ignited, leading to a launchpad abort.

“Karl Henize was pounding on his leg, really mad because he didn’t get to go,” Fullerton recalled.

I turned around to Karl and said, “We don’t want to go, Karl. There’s some­thing wrong out there, you know. " We were worried then: is this going to. . . mess everything up? It did to some extent, but the ground worked overtime, be­cause everything was sequenced by time because it’s an astronomy thing. Wheth­er we’re on the dark side or the light side, all that had to be rewritten. And it all worked out great. We even made up for the fuel we’d had to dump because of the engine failure on the way up [on our second attempt] and eked out an extra day on it. We were scheduled for seven and made it eight.

Once on orbit, the crew followed the established Spacelab pattern of working twenty-four hours a day, in two, twelve-hour shifts. “I anchored my schedule to overlap transitions, so if something came up on one shift, I could learn about it and carry it over to the next shift, hopefully,” Ful­lerton recalled.

But I also had to stagger things so I got on the right shift for entry, so I was in some kind of reasonable shape at the end of the mission. At the beginning, too, we had the red team sleeping right up till launch time so that once we got on orbit, the red team was the first one up, and they’d go for it for twelve hours. So it was all that kind of thing, juggling around so that the right people that had to be alert for launch and entry were. We got into that circadian cycle prior to launch. So the last week they didn’t see the other team; I only saw part of one and part of the other myself.

Compared to his first mission on a two-person demonstration flight, the presence of a larger, seven-member crew this time around had down­sides as well as benefits.

The pressure is higher when you’re commander—the pressure of making sure that not only you, but somebody else, doesn’t throw the wrong switch. With Jack and I, it was just the two of us. He only had to worry about me, and I him. We could double-check each other. With seven people, there are many opportunities for somebody to blow it, not to say instant disaster, but to use too much fuel or to overheat some system or not have the right ones on and blow the chance to get this data. . . . That’s a lot of other people throwing switches, too.

In addition to the responsibility of supervising the crew’s actions, Fuller­ton found many other pressures involved in the role of commander.

During the entry, there was the pressure, [of] it’s your fault if this doesn’t come out right. When you’re in the [pilot’s] right seat, it’s not all your fault. The com­mander bears culpability even if you make a mistake. I’m dwelling on this pres­sure thing because that really is a strong part of the challenge. I mean, you’re really tired after spaceflight. I think you’re tired mostly because you elevate your­self to this mental high level of awareness that you’re maintaining. Even when you’re trying to sleep, you’re worried about this and that. So it’s not like you’re just lollygagging around and having a good time. You’re always thinking about what’s next and mostly clock watching. Flying in orbit is watching a clock. Every­thing’s keyed to time, and so you’re worried about missing something, being late.

The mission featured another historic footnote. In the mid-1980s the “Cola Wars” between competing soda brands Coca-Cola and Pepsi-Cola were in full swing, and on this mission the competition moved into space. Each company attempted to find a way to dispense its beverages in micro­gravity so that astronauts could drink them on the shuttle. Gravity plays an important role in the proper mix of carbonated sodas, and in microgravity the carbonation separates. The goal was to create a device that would en­sure the beverages were properly mixed as they were consumed. Ultimate­ly, however, the unique factors of the near-weightless environment limited the success of both attempts.

Astronaut Don Peterson, who had already left the astronaut corps at the time, recalled an incident during the 51F flight that demonstrated just how many people it took to make the shuttle successful, including a large number of people who never receive any recognition. “I remember, this. . . was one of the flights that Story Musgrave was on, because he and I were friends and I was kind ofwatching. During launch, they got an indication from instrumen­tation that one of the engines on the orbiter was overheating; . . . it was over­pressure or overheat, something, and they shut it down,” Peterson remembered.

They were far enough along in launch that they could still get to orbit. Well, then they got the same indication on the second engine. Now, if you shut down a second engine, you’re into an abort, and that’s a pretty messy operation. There was a young woman [who] looked at that. She was the booster control in the

Mission Control Center and this is happening in real time. You’ve [got] to re­alize, this sucker’s up there burning away and you’ve got people, human be­ings, on board and all that. She looked at that and said, “I don’t believe we’ve got two engine failures on the same flight. That’s highly improbable. I believe we’ve got an instrumentation failure. Don’t shut the engine down. " Now, in hindsight, that was a wonderful decision, but had she been wrong, the back end of the vehicle would have blown out and killed everybody on board and lost a shuttle.

While the woman’s decision saved the mission, she received little recogni­tion for her split-second call, outside of nasa insiders closest to the situation.

Of course, the flight crew and all us people thanked her profusely, and she was recognized, but I don’t think she ever really got any public recognition for that at all. But, I mean, that’s a life-or-death decision under tremendous pressure with. . . human beings and a two-and-a-half-billion-dollar vehicle. And you can’t get under much more pressure than that, and she called it right. What I did to her was terrible. I called her up later. I waited about a month and called her up and said, “I’m Robert Smith, and I’m a reporter with Life magazine. I understand that you’re the woman that saved Story Musgrave’s life." And there was this long, long silence. And finally she said, “Who the hell is this?" Story has a reputation as a lady’s man. So she kind of got a kick out of that, I think.

sts-61A

Crew: Commander Hank Hartsfield, Pilot Steven Nagel, Mission Spe­cialists Bonnie Dunbar, James Buchli, and Guion Bluford, Payload Specialists Reinhard Furrer, Ernst Messerschmid, and Wubbo Ockels Orbiter: Challenger Launched: 30 October 1985 Landed: 6 November 1985 Mission: German Spacelab laboratory flight

The fourth Spacelab flight, 61A, was yet another mission that introduced new elements to the shuttle program by building on its predecessors, par­ticularly the first Spacelab flight. STS-9 had been distinguished from previ­ous shuttle missions in part for carrying the first European crew member, West German astronaut Ulf Merbold. If STS-9 had been a further step to­ward international spaceflight, 6ia was a leap—a mission purchased by and dedicated to the work of a foreign country, West Germany.

Commander Hank Hartsfield recalled explaining to a West German reporter the circumstances of Germany paying to use the German-built Spacelab module. “He said he wants to know how much Germany has to pay the United States to use their Spacelab, because Spacelab was built in Germany. It was built in Bremen. They were very sensitive about it. I think Germany had paid eighty million dollars for that flight. But this report­er was taking a very nationalistic look at it: ‘We built it and now we have to pay to use it.’” Hartsfield explained that what the reporter failed to un­derstand was that, while West Germany had built the modules, it did not own them. The first Spacelab module West Germany built, lmi, was given to the United States in exchange for flying Payload Specialist Ulf Merbold and experiments, and the second, LM2, was purchased by nasa.

According to Hartsfield, working with the Germans in planning the mis­sion was an interesting process, filled with “delicate” negotiations. While he explained that he would stop short of describing nasa’s West German part­ners as “demanding,” they definitely had expectations for what they wanted out of the mission. “They pushed hard to get what they wanted out of the contract that they had signed with the U. S., and they took an approach where they would hang up on words, on what a word meant, in the agreement.”

For example, one of the biggest controversies Hartsfield remembered is what language would be used on the mission.

They wanted to use the German language and talk to the ground crews in Ger­many and speak German. I opposed that for safety reasons. We cant have things going on in which my part of the payload crew can’t understand what they’re getting ready to do. It was clearly up front, the operational language will be English. We fought that one hard. We finally cut a deal that in special cases, where there was real urgencies, that we could have another language used, but before any action is taken, it has to be translated into English so that the com­mander or my other shift operator lead and the payload crew can understand it. . . . There were several times we did use the language during the flight, they asked for German, and it all worked real well. In fact, one case, I think it was national pride. Somebody wanted to talk to Wubbo [Ockels]—he was from the Netherlands—and they wanted to speak Dutch. Somebody insisted they had this urgent thing, and a friend of mine that spoke Dutch said, “You know what the guy wanted to do? He wanted to say, ‘Hello, Wubbo, how’s it going?”’

Hartsfield recalled a conversation with West German mission manager Han – sulrich Steimle, whom he described as “a very interesting fellow” who “liked to philosophize.” During preparation for the mission, the two became friends and would discuss cultural differences between the two countries. “He says, ‘You know, in the United States, when a new policy comes down, the Americans, they look at this and say, ‘Okay, here’s what we’ve got to do,’ and they salute and go do it. He says, ‘In Germany, when a new policy comes down, we study it very carefully, and decide how we can continue to do what we’re doing un­der this new policy.’ I thought, ‘Boy, is that ever true.’ I know some people in the United States that do that, too, you know. I worked with some of them.”

The multinational nature of the crew popped up in various and inter­esting ways, from the larger issues that led to the “delicate negotiations” to more trivial and potentially amusing incidents, Hartsfield recalled.

We were training in Building 5 [at Johnson Space Center], and once we went into quarantine,… we had to use the back door, and they issued keys to us. The keys had two letters and a number on it that identified that particular key. . . . Well, when we issued the keys, Ernst [Messerschmid, one of the German astro­nauts] came to me, and he was pulling my chain, because he’s a wonderful per­sonality, he said, “Is there any significance to the fact that I got a key with ss on it?” [“ss ” was the abbreviation for the Schutzstaffel, an official Nazi paramil­itary organization during World War II.] So we got a big laugh out of it. He did. “No,” I said, “it’s the luck of the draw.”

The mission was the first to be partly directed from outside the Unit­ed States. Germany had built its own control center in Oberphaffenhofen, southwest of Munich. Yet another of the interesting cultural differences came to light during a visit by the crew to the control center during mission preparations. “They had an integrated sim, the first one they’d run, and the whole payload crew was over there working in the simulator and doing this sim to get their controllers up to speed,” Hartsfield recalled.

Well, they were also filming a documentary on how Germany was preparing for this thing, and so there were camera crews walking around in their control center taking pictures, and it got to be lunchtime. Things are different in Ger­many, you know, about drinking. To have wine and beer at lunch is a common thing. Well, in the basement of the control center, like in a lot of German busi­nesses, they’ve got machines like Coke machines, but it’s a beer machine. The flight controllers had all gone down and got a beer, and here is this crew, all of them got a beer, sitting on the console, and eating lunch. I called Hansulrich Steimle, and I said, “Hansulrich, I know how things are here in Germany, but you’re filming for posterity here. ” I said, “If this film goes outside of Germany, some people may not understand your flight controllers drinking on duty. ” Then I became very unpopular, because some of them knew that I had said something, because he made them put their beer away.

The cultural differences were particularly an issue during mission prep­arations for Bonnie Dunbar, the only female member of the crew. As the two nasa mission specialists, she and Guy Bluford were sent over early to begin training in West Germany.

When I showed up there was a lot of discussion about both of us, but, with re­spect to me, they were very concerned that I had been assigned to the flight, be­cause their medical experiment wasn’t intended to include female blood. They thought that would ruin their statistics. . . . I was actually told in front of my face—and I have to first of all qualify that I’ve become very good friends with all these people; but any time you’re at the point of the pathfinder, there’s going to be things happening—I was told that maybe NASA had done this intention­ally to offend the Germans by assigning a woman.

As they started to work through that, it came to light that some of the experiments on the mission, such as the vestibular sled experiment, did not fit her, and the Germans began saying they were going to need for her to be replaced. This was Dunbar’s first flight assignment, and she began to wor­ry that she would lose her seat. “At that time Dr. Joe Kerwin was head of Space and Life Sciences. . . . He actually wrote a memo for the record and to dfvlr [Deutsche Forschungs Versuchsanstat fur Luft und Raumfahrt, the German Aerospace Research Establishment] stating that all equipment should be designed to this percentile spread. So they stood by [me].”

However, Dunbar also had many positive experiences working with the West Germans in preparation for the mission and enjoyed their excitement about it. She began working with the Germans during preparations for the first Spacelab flight, and she said it was fun to see the atmosphere there dur­ing their entry into working with human-rated vehicles.

When I went over there, it was a very exciting time for them. The Bundestag had become involved in this, which is their legislative government. At that time the seat of government in [West] Germany was in Bonn, and were training in Bonn/Cologne. So it wasn’t unusual for our two German astronauts, Ernst Messerschmid and Reinhard Furrer, to have lunch with what we’d call a sena­tor. There was just a lot of interest. The mission manager at that time told me that they hoped to use this mission not only to advance their science and human spaceflight but to inspire a generation of young people in Germany that real­ly hadn’t had inspiration since the war, and they lost the war. So this had very much a political flavor to it, not just a scientific flavor to it.

According to Dunbar, the German astronauts were great to work with, and it was very interesting being around them in West Germany, where they had a respected, heroic status.

Actually, it was fun to talk to them, because in Germany they very much oc­cupied an Original [Mercury] Seven status, and so it was always interesting to hear from their perspective and see what they were doing, how the program was being received. It’s changing now, but there’s still a lot of dichotomy, even in Germany, about the value ofhuman spaceflight, even though they’re a major partner in esa [the European Space Agency] and in the Space Station, because there would be one or two ministers that would vocally try to kill the program, and so we were interested in working with them and helping them.

Hartsfield recalled several highlights of 61A once the preparations were complete and the flight was finally underway. “I would say it was probably the most diverse mission ever flown,” Hartsfield asserted.

We had a black, a woman, two Germans, a Dutchman, and a marine. I mean, how diverse can you get, you know? And there was some funny things hap­pened. We launched on October 30, and of course, October 31 was Halloween. So I took a back off one of the ascent checklists we weren’t going to use anymore, and I drew a face on it, cut out eyeholes, got some string, and I made myself a mask. I took one of the stowage bags and went trick-or-treating back in the lab. Of course, they don’t do Halloween in Germany, or Europe, so they didn’t know what I was up to. I decided not to pull any tricks on them, but I didn’t get much in my bag. But somebody took a picture. One of the guys took a pic­ture of me with that mask on, holding that bag, and somehow that picture got released back in the U. S. About a month after the flight, I got a letter from NASA headquarters. Actually, the letter had come from a congressman who had a complaint from one of his constituents about her tax money being spent to buy toys for astronauts. She was very upset. So it was sent to me to answer, and I had to explain, hey, nothing was done, and it was made in flight from mate­rial we didn’t need anymore. It was just fun. I never heard any more, so I think maybe that satisfied her. She had the notion that we had bought this mask and bag and stuff just to do Halloween.

Another amusing anecdote Hartsfield recalled involved Bluford.

Steve [Nagel] and I were up on the flight deck. All of a sudden, we heard this bang, bang, bang! It sounded like somebody was tearing up the mid-deck. We peeked our heads down in the hole on the side where the bunks were. About that time we saw the bottom bunk come open, and the top one. Bonnie is stick­ing her head out looking up and Ernst is looking down, and all this banging is going [on] in this little bunk. So they slide Guy’s door open to his bunk, and he kind of looks around, “Oh," and he pulls it back shut and goes back to sleep…. Apparently, he had awakened and didn’t know where he was. He had a little claustrophobia or something, and he was completely disoriented, you know. But when he finally saw where he was, he said, “Okay," and he went back to sleep.

Hartsfield also recalled the crew flying through a mysterious cloud of particles during orbit. “Steve was up on the flight deck with Buckley, and they yelled back at me, ‘Hey, Hank, get up on the flight deck.’ And I got up there and looked out the window, and there was these little light things, bouncing off the windscreen. At first I couldn’t tell how fast they were going. Zoom! Zoom! They looked like they were going real fast. And I said, well, it can’t be that fast, and couldn’t be massive, because they aren’t breaking the window or anything. We were down close to Antarctica, at the south­ernmost point of our orbit.”

For the duration of the mission, the crew had no idea what was causing the phenomenon. In fact, according to Hartsfield, answers didn’t come un­til after the landing, when engineers looked at the orbiter’s front windows. “When the solids [separated during launch], it would coat the windscreen. When those little things hit it, it was cleaning the windscreen. Those spots just took the grease right off of it. When we got back, we found out it was water. We had done a water dump on the previous rev. And we didn’t real­ize it, but all of that turned into ice particles. And then we flew through it. It was weird. We were looking at the window, zoom zoom zoom. It would hit the windows and bounce off, and you’re wondering what the hell it was.”

Secret Missions

From its payload-carrying capacity to the wings that provided its substan­tial cross-range, the Space Shuttle was heavily shaped by the role the U. S. Department of Defense played in its origins. After Congress essentially pit­ted nasa’s Skylab program and the air force’s Manned Orbiting Laborato­ry against each other for funding in the late 1960s, nasa decided to try to avoid such problems with its next vehicle by soliciting DoD involvement from the beginning. In several ways, the shuttle’s design and capabilities were influenced by uses the military had in mind for the vehicle.

Until early 1985, however, the military played only a limited role in the shuttle’s use. Beginning as early as STS-4, there had been flights with classi­fied military components, but there had yet to be a dedicated military clas­sified flight. That would change with 51c, in January 1985.

STS-51C

Crew: Commander T. K. Mattingly, Pilot Loren Shriver, Mission Special­ists Ellison Onizuka and James Buchli, Payload Specialist Gary Payton

Orbiter: Challenger

Launched: 24 January 1985

Landed: 27 January 1985

Mission: Launch of classified military intelligence satellite

T. K. Mattingly had dealt with classified elements previously, as com­mander of STS-4. After he returned from that mission, Deke Slayton asked him if he would be interested in staying in the astronaut corps and com­manding the first fully DoD-dedicated classified mission. Slayton told Mattingly that the mission should require only six months of training, which would be a very short turnaround compared to many flights. Mat­tingly recalled, “With all the training and all of the years we put into the program, the idea of turning around and going right away was very ap­

pealing, to get my money back for all that time. . . and so I said, ‘Yeah, I’ll do that.’”

The pilot of 51c, Loren Shriver, said the rest of the crew was chosen to bridge the two worlds involved in the mission. Rounding out the crew were air force colonel Ellison Onizuka, marine corps colonel James Buch – li, and air force major Gary Payton. “We knew that STS-10 [as it was orig­inally called] was going to be DoD,” Shriver noted, “and when the crew was formed, it was all military guys that formed the crew. I think NASA be­lieved that it didn’t have to do that, but I think it also believed that things would probably go a lot smoother if they did. So they named an all-active – duty military crew.”

Although the promised quick turnaround had been the drawing card for the mission for Mattingly, problems with a solid-fuel engine used to deploy a payload on one of the shuttle’s first operational missions caused a delay for the mission, since the plan was for 51c to also use a solid-fuel booster to deploy its classified payload. The flight was grounded for more than a year.

During the delay, Shriver learned that being assigned to a crew could have a downside. Traditionally, being named to a crew had been one of the best things that could happen to astronauts—they knew they were going to fly, they knew what their mission was going to be, and they had some idea of roughly when they would fly. With STS-10, which was renamed 51c during the delay, the crew discovered that sometimes being part of a crew could actually keep you from flying. “I thought, ‘Well, maybe I never will fly.’” Shriver said. “It was the kind of situation where once you were identi­fied as the crew for that mission, then especially this one being a DoD mis­sion, you were kind of linked to it, as long as there was some thought that it was going to happen. And it never did completely go away. It just went kind of inactive for a while and then came back as 51c.”

The classified payload for the mission was reportedly the Magnum satel­lite, a National Security Agency satellite used to monitor military transmis­sions from the Soviet Union and China. While the mission was officially clas­sified, according to news reports at the time, information about the payload and its purpose is available in congressional testimony and technical journals.

Everything about the mission was classified, not just the payload. This included all details about training, astronaut travel, and even the launch date. “I couldn’t go home and tell my wife what we were doing, anything about the mission,” Mattingly said. “Everybody else’s mission, everybody in the world knew exactly what was going on; NASA’s system is so wide open. They could tell their wives about it, their family knew, everybody else in the world knew what was on those missions. We couldn’t talk about anything. We couldn’t say what we were doing, what we had, what we were not do­ing, anything that would imply the launch date, the launch time, the tra­jectory, the inclination, the altitude, anything about what we were doing in training. All that was classified. Couldn’t talk about anything.”

People ask questions all the time, Mattingly said, and they ask even more questions when they know that they can’t know the answer.

Then they just get even more adamant that you should tell them and try to dream up of more tricky ways to get you to say something—the media, of course, being number one in that game. . . . Everybody had an opinion as to what it was, and you’d just say, “Cannot confirm or deny," and that’s all that was nec­essary. . . . It was humorous, I guess, to listen to people out there trying to guess as to what it might be. You’d say, “Okay. Well, just let that churn around out there. I’m going to go do my training and not worry about it." And eventually you don’t think much about it. But it does require you, then, when you do go meet the press or you do go do public presentations, that you have to think a lit­tle bit harder about what you can say and not say.

As a result of the delay, Mattingly and crew spent a substantial amount of time preparing for the mission, in an unusual experience that blended the very different cultures of NASA and the military. “The interesting thing about the classified mission is JSC and the whole NASA team has worked so hard at building a system that insists on clear, timely communication,” Mattingly said. “The business is so complex that we can’t afford to have se­crets. We can’t afford to have people that might not know about something, even if it’s not an anomaly. For something that’s different, something un­usual, we try to make sure that it’s known in case it means something to somebody in this integrated vehicle.”

But for a classified military mission the number of people who could know and talk about the mission details was limited. “I had some apprehensions about could we keep the exchange of information timely and clear in this small community when everybody around us is telling anything they want, and we’re kind of keeping these secrets,” Mattingly explained. “Security was

the challenge of the mission. How do you plan for it? How do you protect things? We went around putting cipher locks on all the training facilities, but then you had to give the code to a thousand people so you could go to work.”

Shriver also recalled concerns that the requirement of classification of el­ements of the mission would interfere with the open communication that was a vital part of safety and mission assurance. “In the nasa system, ev­erything is completely open; . . . everybody is pretty well assured of having the information that they know they need. We were concerned that just the opposite was going to happen, that because of the classification surround­ing the mission, people were going to start keeping secrets from each other and that there was a potential that some important product or piece of in­formation might not get circulated as it should.”

In reality, Shriver opined, nasa and the DoD managed to find ways to make compromises that protected the information that needed to be pro­tected while allowing for sharing of information that needed to be shared.

The mission, yes, was classified. Certain descriptive details about what was go­ing on were always classified, but within that classification shell, so to speak, the system was able to find a way to operate and operate very efficiently, I thought. There were still some hiccoughs here and there about how data got passed back and forth, and who could be around for training and who couldn’t, and that sort of thing. But eventually all that got worked through fairly well, and I think the pathfinding we did on that mission helped some of the subsequent DoD – focused missions to be able to go a little bit smoother.

A special room for storing classified documents and having classified conversations was added to the Astronaut Office. The new ready room in­cluded a classified telephone line. “They said, ‘If certain people need to get hold of you, they’ll call this number,’” Mattingly recalled.

It’s not listed and it’s not in the telephone book or anything. It’s an unlisted num­ber, and this causes less attention. “You’ve got to keep this out of sight, don’t let anybody know you’ve got it, and this is how we’ll talk to you on very sensitive things." So we had a little desk in there, put it in a drawer, and closed it up. In the year we worked on that mission, we spent a lot of hours in this little room because it was the only place we could lay our stuff out; the phone rang once, and yes, they wanted to know if I’d like to buy mci [telephone] service.

Even if, by and large, the mission stakeholders did a good job making sure that the secrecy didn’t prevent needed information from being shared, there was at least one occasion when it went too far, Mattingly recalled. “My sec­retary came in one day, and she was getting used to the idea that there’s a lot of people we deal with that she doesn’t know. . . . She came in to me one day, and she says, ‘You just got an urgent call.’ ‘Okay.’ ‘Joe (or somebody) says call immediately.’ So, okay. ‘Joe who?’ She said, ‘He wouldn’t tell me. He said you’d know.’ We went in our little classified room and said, ‘Does anybody know a Joe?’ We never did figure out who it was, and he never called back.”

Preparing for a classified mission was quite the adjustment because of the differences between NASA and air force security systems and the many rules for how to deal with classified information. “In any bureaucracy we sometimes overdo things, but as much [as] we make fun of these folks, they convinced me that some of the precautions we were taking were, in fact, justified,” Mattingly stated.

I was a bit skeptical, but they showed me some things that at least I bought into. Whenever we traveled, they wanted to keep secret when was the launch time, and they certainly wanted to keep secret what the payload mission was. And to keep the payload mission secret, that meant whenever we went somewhere they wanted us to not make an easy trail when we’d go somewhere. To keep the launch time classified, they wanted us to make all our training as much in the daytime as at night, so that someone observing us wouldn’t be able to figure this out. They nev­er convinced me that anyone cared, but they did convince me that if you watch these signatures you could figure it out, and it is secret because we said it was.

The extra work to preserve secrecy, in Mattingly’s opinion, turned out to have some benefits, but also plenty of downsides to go along with them.

I didn’t mind the idea of flying equal [time in] day and night, because that meant I got to fly more, because I wasn’t about to split the time, we’ll just double it. So that was a good deal. But then they had this idea they wanted us, when­ever we went to a contractor that was associated with the payload or with the people we were working with, they didn’t want us to get in our airplane and fly to that location. They wanted us to file [flightplans] to go to Denver and then refile in flight and divert to a new place so that somebody who was track­ing our flight plans wouldn’t know. And when we’d get there, we could check in using our own names at the motel, but, you know, just Tom, Dick, and Harry. So, just keep a low profile.

Even so, the mission preparations did, at times, demonstrate the dif­ficulty of keeping secrets when too many people know the information.

We went out to Sunnyvale [California], and we were going to a series of class­es out there, and this was supposed to be one of these where you don’t tell any­body where you’re going, don’t tell your family where you’re going to be, just go. But the secretary got a room for us. So we went, landed at one place, went over to another place, landed out there at Ames, had this junky old car that could hardly run. El [Onizuka] was driving, and Loren and Jim [Buchli] and I were crammed in this little tiny thing, and we’re going down the road and looking for a motel. And we didn’t stay in the motel we normally would stay at. They put us up and tell us to go to some other place and they had given us a name. So we went to this other place, and it was very inconvenient and quite a ways out of the way. And as we drive up the road, Buchli looks out the window and he says, “Stop here. " So we pull over, and he says, “Now let’s go over [the security proce­dures] one more time. We made extra stops to make sure that we wouldn’t come here directly, and they cant trace our flight plan. And we didn’t tell our fami­lies, we didn’t tell anybody where we are. And we can’t tell anybody who we’re visiting." He says, “Look at that motel. What does that marquee say?" “Wel­come STS-51C astronauts, "andeverybody’s name is in it, andyou walk in and your pictures are on the wall. Says, “How’s that for security?"

Security lapses such as that proved frustrating for the crew, Mattingly re­called, when they themselves made great efforts to preserve secrecy only to watch the information get out anyway. “Those are dumb things, but they show that we went to extraordinary lengths trying to learn how to do some of these things. And the coup de grace came when, after, you know, ‘I’ll cut my tongue off if I ever tell anybody what this payload is,’ and some air force guy in the Pentagon decides to hold a briefing and tell them, before we launched, after we’d done all these crazy things. God knows how much money we spent on various security precautions and things.”

Even with the things that were revealed, the public face of the classi­fied mission was also unusual, Mattingly said, explaining that the pub­lic affairs people at Mission Control had an interesting time dealing with the media. “For the first time the mocr’s [Mission Operations Control Room] not going to be open for visitors, there’s nothing to say, nothing to do, you know. ‘They launched.’ ‘Yeah, we saw that.’ ‘Oh, they came back.’ ‘That’s good.’”

On launch day, those tuning in heard the usual launch discussions at launch control, but the communications between ground control and the astronauts were not broadcast, as they had been on all previous flights.

Shriver said that the secrecy had personal ramifications for the astro­nauts, who wanted to share with their friends and family their excitement about the flight—in Shriver’s case, his first.

The airforce did not even want the launch date released. They didn’t want the crew member names released. We weren’t going to be able to invite guests for the launch in the beginning. This is your lifetime dream and ambition. You’re finally an astronaut, and you’re going to go fly the Space Shuttle, and you can’t invite anybody to come watch. It was an interesting process. We finally got them talked into letting us invite [people]; I think each one of us could invite thirty people, and then maybe some other car-pass guests who could drive out on the causeway. But trying to decide who, among all of your relatives and your wife’s relatives, are going to be among the thirty who get to come see the launch, well, it’s a career-limiting kind of decision if you make the wrong decision.

All in all, Mattingly said, he considered the mission to be a success and was proud to be a part of it. He said the real contributors were the people who prepared the payloads but that it was an honor to have the opportu­nity to deliver them.

I still cant talk about what the missions were, but I can tell you that I’ve been around a lot of classified stuff and most of it is overclassified by lots. I think at best it’s classified to protect the owners, you know, it’s self-protection. What those programs did are spectacular, they are worth classifying, and when the books are written and somebody finally comes out and tells that chapter, everybody is go­ing to be proud. Now, all the things we did for security didn’t add one bit, not one bit. But the missions were worth doing, really were. The work was done by others, but just to know that you had a chance to participate in something that was that magnificent is really kind of interesting.

STS-5IJ

Crew: Commander Bo Bobko, Pilot Ron Grabe, Mission Specialists David Hilmers and Bob Stewart, Payload Specialist William Pailes

Orbiter: Atlantis

Launched: 3 October 1985

Landed: 7 October 1985

Mission: Deployment of two military communications satellites, first flight of Atlantis

Bo Bobko was assigned to command the next classified military mis­sion, 51j, which flew in October 1985. Like 51c, its crew also had a strong military presence: Pilot Ron Grabe of the U. S. Air Force, David Hilmers, U. S. Marines, Bob Stewart, U. S. Army, and Payload Specialist William Pailes, Air Force.

Despite the mission’s classified nature, Commander Bo Bobko described the mission as “pretty vanilla.” “I mean, we went on time and we landed according to the schedule,” Bobko said. “The fact that it was classified was a pain, but you lived with that. Somebody might be doing an experiment, and he could be working on the experiment out in the open as long as it was away from NASA, but having the experiment associated with that shut­tle flight was the classified part of it. So I couldn’t call a person, because as the commander, if I called them, it would give an indication that that ex­periment was on that shuttle flight.”

The 51J mission was the first flight of Atlantis. Bobko, who had also served on the maiden voyage of Challenger, said Atlantis flew well on its first flight.

NASA has subsequently released the information that the payload for the mission included a pair of Defense Satellite Communications System sat­ellites, part of a constellation of satellites placed in geosynchronous orbit to provide high-volume, secure voice and data communications. The sys­tem was a next-generation upgrade from a network the military originally began launching in 1966.

People and Payloads

In 1983 a new classification of astronauts had emerged, joining pilots and mission specialists: payload specialists. Until this point, being an astronaut was a full-time job, a career choice that people committed to for years. They were given broad training, which prepared them to carry out any variety of mission they might be assigned. Payload specialists, on the other hand, had other careers; flying in space, for them, was not their job, but a job duty. Payload specialists were just that—specialists from organizations, includ­ing universities, companies, and nations, responsible for a shuttle payload, who accompanied that payload during flight and helped with its operation.

On 28 November 1983 the sTs-9 mission, also known as Spacelab I, had launched into space the first payload specialists: Byron Lichtenberg, a biomedical engineer from the Massachusetts Institute ofTechnology, and Ulf Merbold, a West German physicist representing the European Space Agency. Lichtenberg was the first American who was not a career astro­naut to fly in an orbiting U. S. spacecraft, carrying out an experiment in space that he helped design and that he would help analyze and interpret as a member of a research team. Before that, scientists had instructed and trained astronauts on how to do their experiments and astronauts did the work for them.

sts-41D

Crew: Commander Hank Hartsfield, Pilot Michael Coats, Mission Specialists Judy Resnik, Steven Hawley, and Mike Mullane,

Payload Specialist Charles Walker Orbiter: Discovery Launched: 30 August 1984 Landed: 5 September 1984 Mission: Deployment of three satellites

The year after Lichtenberg and Merbold’s flight, the first commercial payload specialist, Charlie Walker, flew as a member of the 41D crew. Com­manded by Hank Hartsfield, 41D deployed three satellites and tested the use of a giant solar wing. Walker was assigned to the flight to run the Con­tinuous Flow Electrophoresis System (cfes), an apparatus from the Mc­Donnell Douglas Corporation, for which Walker was a test engineer. The system used electrophoresis, which is the process of separating and purify­ing biological cells and proteins.

“What they were producing with that was erythropoietin,” Hartsfield recalled. “It’s a hormone that stimulates the production of red blood cells. Ortho Pharmaceuticals was the primary contractor with MacDac to build this thing. And cfes was kind of a test version of it. . . . The idea was, which was a good one, say you were going to have planned surgery, they could inject that hormone into you prior to the surgery, some time peri­od, I don’t know how long it would take, but your body would produce more red blood cells, and then you wouldn’t need transfusions. So it was a good idea.”

Prior to being assigned a flight of his own, Walker was responsible for training NASA astronaut crews in the operation of the cfes payload on the STS-4, sts-6, STS-7, and sts-8 shuttle flights during 1982 and 1983. He flew with the cfes equipment as a crew member on 41D, 51D, and 61B.

The initial agreement between McDonnell Douglas and NASA called for six proof-of-concept flights of the cfes device, which was originally to be flown aboard Spacelab. But according to Walker, slips in the launch of Spacelab caused McDonnell Douglas and NASA to renegotiate for six flights on board the Space Shuttle mid-deck.

The agreement was that NASA would provide the launch and McDon­nell Douglas would provide the equipment and testing processes. “nasa, in the body of the Marshall Space Flight Center Materials Lab folks, had the opportunity, at no expense to them other than the preparation of samples and then the collection and the analysis of it later in their own laboratories upon return home, to use a device produced at the expense of the private sector for private-sector research, but, again, allowing NASA the right to use it for up to a third of the time in orbit in exchange for the opportunity to have it there in orbit aboard shuttle,” Walker explained.

After the first flight of the cfes device, on STS-4, Walker said McDon­nell Douglas felt that it had demonstrated enough success in the proof of concept test to ask to fly its own astronaut to run its device. “From our standpoint, we had proven that we could predict adequately for produc­tion processing what we needed to know,” Walker said.

We briefed on that, and we advised the Space Shuttle program management what we wanted to do for the next flight; got that approved through appropri­ate processes, and at the same time—I can remember, I was in a meeting in which Jim Rose [McDonnell Douglas manager] and I briefed Glynn Lunney [shuttleprogram manager] in Glynn’s office in Building 1 [at jsc], and Jim told me, going down, he said, “I just want to tell you, as we walk into this meeting if I get an indication from Glynn that he’s happy with the results, too, from the NASA side, I’m going to ask for a payload specialist opportunity. ” He said, ”Are you okay with that?” And I said, “You know I’m okay with that. ”

That was exactly what happened, Walker recalled. Lunney indicated that nasa was pleased with the results, so Rose pressed ahead.

Jim Rose said, “We want to ask for the opportunity to negotiate for one more thing.” And basically it went something like, “You know, Glynn, the astro­nauts that were training, Hank’s a great individual, obviously a great test pi­lot, a good engineer, but Hank doesn’t know this electrophoresis stuff and the other astronauts, mission specialists that we’re going to train, they’re going to be able just to spend a little bit of their time working with our device. They’ve got lots of other things to do. That’s the mandate for the mission specialist. We really would learn the most we possibly can and more than we can do with a mission specialist if we get the opportunity to have a payload specialist devoted specifically to the electrophoresis device and its research and development activ­ities during a flight. ”

As I remember it, Glynn chewed on his cigar a little bit—that’s when you [could] smoke in the office—chewed on his cigar a little bit and said something like, “Well, we’ve been wanting to move into this payload specialist thing, so if you’ve got somebody that is qualified, can meet all the astronaut selection crite­ria, put in the application. Let’s do it. ”. . . I think Glynn said something like, “Do you have somebody in mind?” Jim turned to me and looked at me, and Jim said something like, “You’re looking at him.” Glynn said, “You, huh?” And I said, “Yes, it’s me. I’ll be the man. ”

Walker was added to the crew for 41D in late May 1983. Within a few days of being added, Walker, who was residing in St. Louis at the time, went down to Johnson Space Center to officially meet the crew, some of whom he knew from the training he had provided for previous missions, and to begin working out a training schedule. The training syllabus was based loosely on the training developed for science payload specialists on Spacelab missions and was to be a foundation for future commercial pay­load specialists. The training syllabus was the result of ongoing negotiations between Walker, his employers, and various stakeholders at nasa. Walk­er enjoyed the training and felt it was important. Unlike other astronauts, however, for whom mission training was their job, training took Walker away from his day-job duties in St. Louis, so while his employers wanted him trained, they also wanted him available there when it wasn’t necessary for him to be in Houston. NASA also wanted to make sure he received all the training needed to fly safely but didn’t want to invest unnecessary time and resources that were needed for career astronauts.

Walker explained that the payload specialist training was a shorter, con­densed version of the career astronaut training. Initially, he said, that con­densed training still involved at least an overview of a wide variety of sys­tems. “Hank Hartsfield had me operating the remote manipulator system, the rms, in the trainer, and that went on for a few weeks. I was training with the crew. I was working the rms in the simulator, and I knew the sys­tem, I knew how to work it, even though I was not in the flight plan to de­ploy any of the satellites or to have to use the rms, as might conditionally be the case. I don’t think on [41D] we had any required use of the rms, but we had as a contingent the operation of it.”

Later, however, management came back and decided that Walker didn’t need to be trained on equipment he wasn’t going to use. There was no need, they argued, to invest valuable time and resources training payload specialists on things they wouldn’t be doing. Walker said that regardless of the actual need for the additional training, he felt like it had substan­tial benefits for the mission. “My comments were that I think this is a good thing,” he said. “Let the payload specialist do some of this, too. He or she is going to feel like more of a cohesive part of the crew. It’s just a good psychological thing, even though you don’t need their hands to es­pecially do that.”

The most important and most time-consuming part of preparing for a mission was for payload specialists to gel with the crew, Walker said. “In other words, the crew’s getting to know me, and me getting to know my fellow crewmates for each flight, so that we knew, at least to a significant degree, each other’s characteristics, and we could work together and feel good working together and flying together as a team.”

Walker spent, on average, about two weeks each month training at jsc during the nine months leading up to the flight. He said Hartsfield wanted to integrate him with the crew as closely possible, and in addition to offi­cial training sessions, the commander included him in the occasional social event as well. “I was invited to more than one dinner or activity at Hank’s house and some of the other homes of the astronauts—the crew as well as others—but it wasn’t as close a relationship as was the case between the career astronauts and families down there that were obviously living and working at each other’s elbows day in and day out.”

Hartsfield said he felt so strongly that Walker should be an integral part of the crew that he requested Walker’s name be on the patch circle with the rest of the crew. “[Payload specialists’] flight assignments changed a lot,” Hartsfield said. “Some of the flights had as many as three different peo­ple assigned at one time or another, and they had to keep changing their patches. So to save money, they put a ribbon at the bottom with the pay­load specialist on it, so they wouldn’t have to change the whole patch. But when Charlie flew, I had sold George Abbey on this. ‘He’s part of the crew, you know. Put his name on the patch with the rest of us.’”

Admittedly prejudiced, Hartsfield described the 41D crew as one of the best crews ever put together. “As the commander, I just sort of had to stay out of their way,” he said. “I was reminded of that two-billed hat, you know, that says, ‘I’m their leader. Which way did they go?’”

The addition of payload specialists brought a new dynamic to space­flight and the astronaut corps. Walker’s crewmate Mike Mullane said sev­eral astronauts, including himself, had viewed payload specialists as out­siders and as competitors for flight assignments. “There was some friction there, I think, that we felt like, ‘Hey, why isn’t a mission specialist doing this experiment,’” said Mullane. “But I’m mature enough now, and par­ticularly after you get your one mission under the belt, you become a little more tolerant of the outsiders.”

People and Payloads

28. The crew of STS-4ID. Courtesy NASA.

Walker said there were only a few individuals in the agency from whom he got the impression that he and the other payload specialists were con­sidered outsiders. “I was there as a working passenger. I wasn’t a full-fledged crew member, and I knew that going in, and I took no real exception to that,” Walker said. “Occasionally there were circumstances in which it was made clear to me that ‘You’re not one of us. You’re along for the ride, and you’ve got a job to do.’ But it was only a few individuals, some in the As­tronaut Office, others outside the Astronaut Office, from whom I got that impression.”

One clear distinction between the career astronauts and the payload spe­cialists was the locations of their offices. At Johnson Space Center, the payload specialist office was in Building 39, and the Astronaut Office was in Build­ing 4. According to Walker, “It was made clear to us from the beginning that they didn’t expect to see us over on the fourth floor in Building 4 except for scheduled meetings. We were just outsiders who would become crew mem­bers for a short period of time and would train mostly on our own, but when there was necessary crew combined training, certainly we would be there.”

Walker said he was fine with that arrangement and was just grateful for the opportunity, both professionally and personally. “There was no belliger­ence, really, expressed openly, and no offense on my part taken,” Walker said.

I really saw my role and my place in this, this was a great adventure, and more than an adventure, it’s a great challenge, both to people as well as to technical systems. I think I know my limitations, and I know that I’m not nearly as quali­fied to make critical and rapid decisions in some of these flight environment cir­cumstances, as the men and women that have been selected by the agency through grueling processes, to do just exactly that. . . . I was getting a great opportunity, I felt, both for the company that was my employer, for the commercial as well as the prospective societal benefits from the work that we were proposing to do through and with the Space Shuttle. And certainly, certainly a tremendous per­sonal opportunity for me, and I was just happy to be there.

The crew went out to launch on 26 June 1984, and everything was going smoothly until the clock reached four seconds. “The engines had already start­ed to come up, and then they just shut down,” Hartsfield said. “We looked at the countdown clock on the onboard computer display, and it was stopped at four seconds. We were really checking to see if there was anything out of

the ordinary. We were going to make sure that things were still okay. There was a good moment of tension there, and Hawley broke the tension. As soon as we looked at everything and everything was okay, Steve said, ‘Gee, I kind of thought we’d be a little higher at meco [main engine cutoff].’”

As the Launch Control Center was trying to figure out the problem, a hy­drogen fire was noticed on the pad. “The trouble with a hydrogen fire is you can’t see it,” said Hartsfield. “Hydrogen and oxygen burn clear, and you can just see some heat ripples when you’re looking through it, but you can’t see the flame. I think one of the sensors picked it up, a uv sensor, which can see it.” When the fire was discovered, there was talk about having the crew leave the shuttle via the slide wire Emergency Egress System. No one had ever ridden the slide wire, Hartsfield said, and flight controllers were afraid to tell the crew to do it. “That bothered a lot of us in that they were concerned enough about the fire that they really wanted us to do an emergency egress from the pad area, but since the slide wire had never been ridden by a real live person—they’d thrown sandbags in it and let it go down—they were afraid to use it, which was a bad situation, really.”

Hawley remembered the crew members talking about whether or not they should get out of the orbiter and use the Emergency Egress System baskets. “I remember thinking, well, fire’s not too bad because then you’re sitting inside this structure that’s designed to take several thousand degrees during reentry. It’s well insulated. Then I got to thinking, on the other hand, it’s attached to millions of gallons of rocket fuel, so maybe that’s not so good. But eventually they came and got us.”

As a result of the pad abort, NASA revamped and tested the procedures for the Emergency Egress System and implemented new training for abort­ing and recycling launches.

Steve Hawley said Mullane was very concerned after the fire that the flight was going to be canceled. “I really didn’t think they would do that, but I remember him being very concerned about that, probably more con­cerned than [about] the incident itself. He was concerned about the effect it would have on his flight assignment.”

While the flight did get delayed for two months and its payloads were changed as a result, the mission did not get canceled entirely. Mullane said he found scrubs personally and emotionally draining. “There is nothing that is more exhausting than being pulled out of that cockpit and knowing you have to do it tomorrow,” he said. “It is the most emotionally draining expe­rience I ever had in my life of actually flying on the shuttle. I will admit that it is terrifying to launch. Once you get up there, it’s relaxing, but launch, it’s terrifying. And people assume that it gets easier. I tell people, no, it doesn’t. I was terrified my first launch. I was terrified my second launch. I was terri­fied my third launch. And if I flew a hundred, I’d be terrified on a hundred.”

Mullane said that before every launch he felt like he faced the possibil­ity of death, as if he were preparing to die. “I know it’s ridiculous to think you can predict your death,” Mullane said.

You could get in an auto accident driving out to get in the T-38, and that’s your death, and here you are thinking it’s going to be on a shuttle. But I certainly pre­pared for death in ways, in a formal way. I served in Vietnam, and there was certainly a sense of you might not come back from that. And I said my goodbyes to my parents and to my wife and young kids when I did that, but this time it was different because it’s such a discrete event. It’s not like in combat where in some missions you go off and fly and never see any enemy antiaircraft fire or anything. But this one you knew that it was going to be a very dangerous thing. And as a result, twenty-four hours before launch, you go to that beach house and you say goodbye to your family, to the wife, at least. That is incredibly emotion­al and draining, because the wife knows that it could be the last time she’s ever going to see you, and you know it’s the last time you might ever see her.

Weeks before launch, crew members and their families choose a fam­ily escort to help families with launch details and to be with them during launch. The family escort stands next to the family on top of the Launch Control Center during launch. Part of that role is simply helping the fam­ily get to where they need to be, making sure everything goes smoothly. However, in addition, the family escorts serve as casualty assistance officers in the event something unexpected happens.

Mullane recalled that his wife commented to him, “‘What I’m picking isn’t a family escort; it’s an escort into widowhood.’ You have this buildup, this incredible emotional investment in these launches that just ticks with that clock. Picking the astronaut escort. The goodbye on the beach house, at the beach house, that lonely beach out there. And now to go and get into the cockpit. Like I said, I thought a lot about death. I mean, I felt this was the most dangerous thing I would ever do in my life was ride this shuttle.”

Mullane opined that it was a mistake on NASA’s part to build the shuttle without an escape system. “I don’t know what the thought process was to think that we could build this rocket and not need an escape system, but it was the first high-performance vehicle I was ever going to fly on with no escape system,” said the three-time mission specialist.

If something went wrong, you were dead. So that was the sense of death that kind of rode along with you as you’re driving, preparing for this mission and driving out to the launchpad. You know it’s the most dangerous thing that you’ve ever done in your life. And to get strapped in and be waiting for that launch, and man, I’ll tell you, your heart is in your throat. I mean, after a launch abort, I swear, you could take a gun and point it right at somebody’s forehead, and they’re not even going to blink, because they don’t have any adrenaline left in them; it’s all been used up. To be strapped in out there and then to be told, “Oh, the weather’s bad. We got a mechanical problem, ” and to be pulled out of the cockpit, and now it’s all going to start over. Twenty-four hours you go back, you’re exhausted, you go back, have a shower, meet your wife, say goodbye again, and then start the process all over the next day. And you do that two or three times in a row, and you’re ready for the funny farm. It really is a very emotionally draining thing.

The 41D mission had a total of three scrubs—two in June and one more in August—and finally lifted off on 30 August 1984.

Once the crew was finally in orbit, Steve Hawley, for whom 41D was his first flight, said it took several days to adjust to microgravity, but the team didn’t have several days before starting to work. “It’s interesting because what we’ve always done. . . is plan the mission so that the most important things happen first,” Hawley said.

That goes back to the days when we’d not flown the shuttle before and every­body was concerned that it was going to fall out of the sky, and so if you got up there, you needed to get rid of the satellite or whatever it was right away, so that when a problem happened, you’d have the mission accomplished. But the shut­tle is very reliable, and so what you end up doing is doing the most important, most challenging, most difficult tasks when the crew is the least prepared to do it, because they’re inefficient and they haven’t adapted yet. . . . Back in those days, we were launching satellites five hours after we got on orbit, and we were still trying to figure out how to stay right side up.

The 41D mission deployed three satellites: two Payload Assist Modules and a syncom for the navy. It was the first time three satellites were launched on one flight. The mission also performed a demonstration of the Office ofApplication and Space Technology solar wing, referred to as oast-1. The 102-foot-tall, 13-foot-wide wing carried different types of solar cells. It dem­onstrated the use of large, lightweight solar arrays for future use in build­ing large facilities in space, such as a space station.

As part of the demonstration, oast-1 was extended to its full height several times, stretching out of a canister mounted on a truss in the pay­load bay. “When fully extended, it was 102 feet tall, and really spectacular to look at,” Hartsfield said.

The array did not have very many actual solar cells; instead, it was pri­marily a test structure to see how well the truss would extend. It featured three linear rods with cross-rods and cables, such that the rods were in com­pression and the cables were in tension. The structure collapsed into a cy­lindrical canister for launch. NASA engineers had predicted how rigid the structure would be based on models, and the orbital experiment would give them the opportunity to validate those predictions and models.

“Surprisingly, once the thing is deployed, it’s fairly rigid,” Hartsfield said. “What was interesting was the array was an order of magnitude stiffer than the engineers had predicted, which was a big surprise to them. In fact, by the time we got ready to fire the second set of firings, which was supposed to in­crease the motion, the thing had almost stopped completely, it was so stiff.” Walker’s Continuous Flow Electrophoresis System worked as planned, but postflight analysis showed that the samples had some biological con­tamination. “In other words, a little bit of bacteria had gotten into some of the fluids during preparation before flight, and the bacteria had grown dur­ing flight and contaminated what we intended to have as . . . biologically pure, uncontaminated by extraneous bacteria,” Walker said. “So the work that I had done had been, so to speak, technically productive. We learned new procedures. We validated the procedures. But the veracity of the bio­logical sample itself for the medical testing that we were going to do post­flight turned out to be a problem, turned out to be bad. So we were not a complete success in terms of our mission accomplishment because of that.” One of the more memorable episodes on the flight was the infamous “peecicle.” During flight, the crew had a problem with an icicle forming around the nozzle where they dumped wastewater, primarily urine and con­densation from the orbiter’s humidity control. There was a lot of concern about the icicle because when the orbiter started reentry the frozen water was in just about the right place to break off and hit the Orbital Maneuver­ing System, Hartsfield said. “If you hit the oms pod and broke those tiles, that’s a real high-heat area right on the front of that pod, you could burn through. And if you burned through, that’s where the propellant is for the oms engines, and that’s not a good thing to have happen.”

Hawley recalled that the ground called up and had the crew test the wastewater dump. “I think we didn’t know anything was unusual initial­ly,” Hawley said.

I think maybe the ground called us and told us to terminate the supply water dump because they had seen some temperature funnies. So we did, and then sometime later, I guess they got curious enough to use the cameras on the robot arm to see what was there. So we set the arm up, and yes, you could see this ici­cle there. For whatever reason subsequent to that, they decided that we ought to try a waste dump and watch it with the camera on the arm, and the icicle was still there. I remember, as we were doing it, watching the second icicle form. So we ended up stopping that dump, and now here we are with this icicle.

The ground crew started working to find possible solutions, one of which was to turn that side of the orbiter toward the sun and let the icicle melt. “After about three days we were convinced that the ice was not going to sublime off the orbiter,” Hartsfield said. “It reduced in size somewhat, but it was still there. I had people ask me, ‘Gee whiz, you got it right in the sun, why didn’t it melt?’ I said, ‘The same reason snow and ice don’t melt on a mountain. It’s in direct sunlight, but it doesn’t absorb much heat. It reflects most of it.’ That was the same thing as this icicle. It wasn’t going anywhere.”

The next option was to send astronauts on a spacewalk to break off the ice. Hawley and Mullane had trained as contingency eva crew members and were selected for the eva, if there was to be one. “I remember Mike was thrilled,” Hawley said,

because he was going to get to do a spacewalk, and I’m sitting there going, “This is not a good idea. I don’t know how in the world were going to get to it. ” I mean, it’s down on the side of the orbiter aft of the hatch, and there’s no trans­lation path down there. I guess they were talking about taking the cfes unit apart, using some of the poles that the cfes was constructed with to maybe grab one of us by the boots and hang him over the side and have him knock it off. That all sounded like a bad plan to me.

It was decided not to try an eva but to use the robotic arm to knock the ice off instead. “I remember thinking, ‘Yeah, it’s a good plan,’” Hawley re­called, “and Mike was thinking, ‘Oh no, I’m not going to get to do an eva.’”

While mission controllers were trying to resolve how to get rid of the icicle, the crew faced another, more immediate issue inside the orbiter. To avoid making the icicle larger, the decision was made that the crew would not be able to dump the waste tank again. While there was still some room in the tank, calculations revealed that the condensation that would be col­lected during the rest of the mission would fill that volume. “What that meant practically to us,” Hawley explained, “was that we couldn’t use the toilet anymore, because there was no room in the waste tank for the liq­uid waste.”

The crew members collected their waste in plastic bags and stored the waste-filled bags on board. Walker said some of the bags were left over from the Apollo program. “I’m kind of an amateur historian,” Walker said, “so I felt a little bad at peeing in these historic bags, but we had to do what we had to do.”

In retrospect, Hartsfield said, the incident is funny, but it wasn’t funny at the time. “The problem was that in zero g, Newton’s third law is very apparent to you. If you just try to use a bag, when the urine hit the bot­tom of the bag, it turned around and came right back out, because there’s no gravity to keep it there. Didn’t take long to figure that wasn’t going to work.” The astronauts stuffed the bags with dirty underwear, socks, towels, and washcloths to absorb the urine.

Hartsfield decreed that the only female on the crew would continue to use the shuttle bathroom. “Judy, as you can imagine, had a hard time with the bag, so we had a little room in there. I said, ‘I don’t care what the ground says, you use the bathroom. The rest of us will do the bag trick.’”

The situation was messy, with all of the bags being stored in the waste stor­age tank under the floor. Hartsfield said there was at least one instance where a crew member was stuffing a urine bag into the tank and the bag ruptured.

Twice, Hartsfield said, Flight Director Randy Stone asked management if the crew could convert a water tank to a waste tank. It would have been an easy conversion, Hartsfield said, but at the time there was great concern about turning the orbiters around quickly for the next flights and the re­sponse was that using the water tank as a waste tank would add a week to the process of getting the orbiter ready to fly again.

“I sometimes think I made a mistake,” said Hartsfield.

I probably should have called for a private med conference and told Flight, “Hey, we’ve got a real problem up here.". . . I talked to the guy that headed that room up when we got back, and he apologized. They later found that it wouldn’t have impacted the flow at all. I said, “Joe, you just don’t know what we’re going through up there. " “Well, you should have told somebody. " “I don’t want to put that on the loop. " I mean, in fact, Gerry Griffin, the center direc­tor, when we got back, he expressed his thanks for not putting that on the open. The media would have had a ball with that.

[Initially] we were hoping to stay another day on orbit, because we had enough fuel to do it, but this was not a very good situation. By the time day six came, we were ready to come home.

For the second time, on board the shuttle was an imax camera. Haw­ley recalled that the imax camera pulled film so fast that in zero g it would torque the user like a gyroscope. The camera had a belt drive with a belt guard, but for this flight, it was decided the belt guard wasn’t needed. “I don’t know if we were trying to save weight or what, but we decided we didn’t need this belt guard.” Hawley said. “I’m up there doing something, and all of a sudden I hear this blood-curdling scream. I go floating upstairs to see what had happened, and Judy had gotten her hair caught in this belt for this imax camera, and there was film and hair all over the orbiter. It jammed the camera and the camera blew the circuit breaker that it was plugged into.”

Mullane said he, Resnik, and Hawley were filming the syncom launch when the incident happened. Mullane and Mike Coats cut Resnik’s hair to free her from the camera, and Coats then spent hours picking hair out of the camera gears in order to get the camera working again. The crew dealt with the problem on their own, without reporting it to the ground, con­cerned that if the public found out, the incident would provide fodder for those critical of NASA’s flying female astronauts.

People and Payloads

29. Judy Resnik with several cameras floating around her, including the imax camera in which her hair got tangled. Courtesy nasa.

Throughout the mission, each crew member went about his or her as­signed tasks with very few coordinated crew activities, Hartsfield said. As a result, he made sure that they ate dinner together every night. “You get a quick breakfast snack, and the first thing you know, you’re off on your daily do list,” Hartsfield said, describing a typical day during the mission. “You eat lunch, normally, on the run where you’ve got a lull in your activ­ities. But I had decreed that the evening meal we were all going to eat to­gether. I want one time for the crew to just get together and just chat and have a little fun and say, ‘Okay, where are we? What have we got to do to­morrow?’ and talk about things.”

One night during dinner as a crew, a rather strange thing happened. “We’d prepared our meals, and we were all floating around, holding down on the mid-deck, and all of sudden we heard this knocking noise, like somebody wanted in,” Hartsfield recalled.

It sounded like knocking. Holy crap, what is that? And then we had a traffic jam trying to get through the bulkhead to get up to the flight deck, because it was coming from up that way somewhere. So we got up there, and we were on the night side of the Earth, it was pitch black out there. Steve flipped on the pay­load bay lights. You know those housings take like five minutes. And we said,

“God, what is that!?" We could hear it, whatever it was, was on the starboard side. Steve was the first one to see it. He looked at the gimbal angles on the Ku- Band antenna, and it was banging back and forth. It was something where it was oscillating back and forth. He hit the power switch and turned it off and that did it. And we went, whew. And we told the ground later what had hap­pened, and it never did it anymore, whatever it did. Apparently it got into a range where it kept trying to swap or do something. I never did find out exactly what caused it, but it sure got our attention. Some alien wants in.

The mission lasted six days, and then it was time to come home. Walker said the reentry and landing on 41D was an emotional experience, drawing to a close what he thought at the time was a once-in-a-lifetime experience.

At that point I didn’t know I was going to have any further flights. I thought that was probably it, and it was such an extraordinary experience, and now it was, for sure, over with. I came to sense a real defining moment, a physically and emotionally defining moment. This experience, this great thing called space­flight, . . . probably above everything else, it’s based upon velocity. It’s putting people and machinery at high speed at the right velocity, the right altitude, the right speed, around the Earth till you keep going, and you’re working in this high-velocity environment that we call orbital flight. When you want to come home, you just take out some of that velocity with some rocket energy again, and use the Earth’s atmosphere to slow you down the rest of the way until you come gliding in and lose the last part of the velocity by applying brakes on the runway until you come to a stop.

So I noted in my own mind two definitive points here that really, without debate, start and end this great experience. One is the high-energy event that we call launch, straight up when the rockets start; to the landing and wheels stop on the runway horizontal, and the brakes have taken hold, and the energy is gone, and the spaceship literally rolls to a stop.

With the end of the mission came the completion of the first flight of a commercial payload specialist and an opportunity to evaluate how the idea actually worked in reality. Hawley commented on how well the crew worked together and how well the crew got along with Walker. “It’s more impor­tant who you fly with than what your mission is, and we really had a good time,” Hawley said. “We all got along well. I thought we all had respect

for each other’s capabilities, and it was just a good mix. . . . Charlie was a good guy. He fit in very well. We enjoyed having him as part of the crew.” The flight marked the beginning of the process, over time, of the soft­ening of hard lines between the career astronauts and the payload special­ists. A major milestone, Walker said, was the decision to move the payload specialists into office space with the career astronauts. “Even while I was training for 61В, I had office space. It was, oh, by the way, catch it as you can, but you got office space over on the fourth floor, Building 4. You need a place to sit and work when you’re in town, come on over. Finally they moved the ps Office out of Building 39 over to Building 4 in that time pe­riod just before Challenger was lost.”

The integration of noncareer astronauts became even more complicated as nasa implemented plans to fly even more types of people on the Space Shuttle—academic and industrial payload specialists, U. S. politicians, in­ternational payload specialists, and the first “Teacher in Space” and “Jour­nalist in Space.”

“I think the clearest example as an indicator of how things transformed was to follow the Teacher in Space activity, because originally the Teacher in Space was to be a spaceflight participant/payload specialist, and I wit­nessed a lot of slicing and dicing of just what do you call Christa McAu – liffe,” Walker recalled. “Is she an astronaut? Well, most people at the time at jsc and certainly in the Astronaut Office were, ‘No, she is not an as­tronaut. We were selected by nasa to be astronauts. We’re the astronauts. She’s a payload specialist.’”

STS-51D

Crew: Commander Bo Bobko, Pilot Don Williams, Mission Specialists Rhea Seddon, Jeffrey Hoffman, and David Griggs, Payload Special­ists Charlie Walker and Senator Jake Garn Orbiter: Discovery Launched: 12 April 1985 Landed: 19 April 1985 Mission: Deployment of two satellites

Despite feeling like his first flight would be a once-in-a-lifetime experience, eight months later Charlie Walker was back in space, this time on 51D. The mission deployed two satellites and carried into space several science experi­

ments and yet another payload specialist. This time, in addition to Walker, on the crew was Jake Garn, a U. S. senator from Utah and the first elected official to fly aboard the Space Shuttle.

Garn was added to the crew about two to three months before launch, recalled Commander Bo Bobko. “George Abbey said to me one day, he said, ‘What sort of training program would you have if you had a new pas­senger that was only going to have eight or twelve weeks?’” Bobko recalled. “I said, ‘Why are you asking me that question?’ He said, ‘Because you’ve got a new passenger, and you’ve only got—,’ I don’t know, ten or twelve weeks to flight.”

Walker said Garn had been lobbying for some time with the NASA ad­ministrator to get a chance to make a flight on the Space Shuttle. Garn was chairman of a NASA oversight body within the Appropriations Committee of the U. S. Senate. “Just part of his job; he needed to do it,” Walker said.

Of course, you look at Senator Garns history, and at that point he had some ten thousand hours logged in I don’t know how many different kinds ofaircraft, having learned to fly as a naval aviator, and had gone to the airforce when the navy tried to take his ticket away from him and wouldn’t let him fly again. . .. Jake was very aviation oriented and certainly enamored with the agency’s activi­ties and just wanted to take the opportunity if one could be found. So his lobby­ing paid off and he got the chance to fly. He was still in the Senate and would take the opportunity on weekends to come train down here; would take congres­sional recesses, and instead of going back to his home state, to Utah, he’d come down here to jsc. So he worked his training in and around Senate schedules.

Walker recalled hearing some negative talk around the Astronaut Office after Garn was added to the crew. “I do remember that there was at least hall talk around the Astronaut Office of, ‘Oh, my gosh, now what’s hap­pening here to us? What have we got to put up with now?’” Walker said.

But Jake, from my experience, and here is an outsider talking about another outsider, but I think Jake accommodated himself extraordinarily well in the circumstances. . . . What I saw was a Jake Garn that literally opened himself up to, “Hey, I know my place. I’m just a participant. Just tell me what to do, and I’ll be there when I need to be there, and I’ll do what I’ll [need to] do, and I’ll shut up when I need to shut up, ” And he did, so I think he worked out ex­traordinarily well, and quite frankly, I think the U. S. space program, NASA, has benefited a lot from both his experience and his firsthand relation of NASA and the program back on Capitol Hill. As a firsthand participant in the program, he brought tremendous credibility back to Capitol Hill, and that’s helped a lot. He’s always been a friend of the agency and its programs.

Bobko lauded Garn for knowing what it meant to be part of a crew. “I’d call him up and I’d say, ‘Jake, we need you down here.’ And he’d say, ‘Yes, sir,’ and he’d be down the next day for the sim,” Bobko said.

Garn’s only problem, added Bobko, was that he got very sick on orbit.

He was doing some of these medical experiments, and they find that one of the things that happens is that on orbit, if you get sick, your alimentary canal, your digestive system, seems to close down. So what they had were little microphones on a belt that Jake had strapped to him to see if they could detect the bowel sounds. So the story is—and I haven’t heard it myself-—they had me on the mi­crophone saying to Jake, “Jake, you’ve got to get upstairs and let them see you on TV. Otherwise, they’ll think you died and I threw you overboard. ”

According to Walker, he and Garn were the guinea pigs for quite a few of the experiments on the flight. One of the experiments was the first flight of a U. S. echocardiograph device. “Rhea [Seddon] was going to do echo­cardiography of the hearts of I think at least three of the crew members, and of course, Jake and I were the obvious subjects,” said Walker. “We re­ally didn’t have much of a choice in whether we were going to be subjects or not. ‘You’re a payload specialist; you’re going to be a subject.’”

While in the crew quarters prior to launch, Walker said, Garn was ask­ing him the typical rookie questions about what it’s going to be like and what to expect.

He says, “You’ve done this before. Tell me. Give me the real inside scoop. What’s this going to feel like? What’s it going to be like?” “It’s going to be great, Jake. It’s just going to be great. Just stay calm and enjoy it. ”. . . We got into orbit, and I can remember there was the usual over-the-intercom exuberant pronounce­ments, “Yee-ha, were in space, ” yadda, yadda, yadda.

I can remember shaking hands, my right hand probably with Jake’s left, gloves on, and “We’re here, ” and then Jake and I both kind of look at each other, and we’re both beginning to feel weightlessness.

The crew was originally assigned as 51E, but that mission was canceled and the payloads were remanifested as 51D. The mission deployed a commu­nications satellite and syncom iv-3 (also called leasat-3). But the space­craft sequencer on the syncom iv-3 failed to initiate after deployment. The mission was extended two days to make certain the sequencer start lever was in the proper position. Griggs and Hoffman performed a spacewalk to attach flyswatter-like devices to the remote manipulator system. Rhea Sed – don then used the shuttle’s remote manipulator system to engage the satel­lite lever, but the postdeployment sequence still did not start.

“Once it became clear that there was a problem, we got a little depressed,” Walker said. “You train for these things to happen. You know they’re real­ly important. Here’s hundreds of millions of dollars’ worth of satellite out there. Your flight’s not that inexpensive, of course, to send people into space. So a lot of effort has gone into getting this thing up there and to launch­ing it and to turning it on and having it operate, in this case, for the Unit­ed States Navy. And here it didn’t happen, so you’re like, ‘Oh, my gosh.’” The crew immediately began to think in terms of contingencies. Walker recalled a strong awareness of the nearness of the satellite. Despite the fact that it had failed, it was still there, floating not that far away. It was still reachable and could potentially be repaired or recovered.

Within a few days the ground came up with the suspected culprit—a mechanical switch, about the size of a finger, on the side of the satellite was supposed to have switched the timer on. The thought was that maybe that switch just needed to be flipped into the right position. If the shut­tle could rendezvous with the satellite, all that would be needed would be some way to flip the switch.

The ground crew instructed the flight crew to fashion two tools that Walker referred to as the “flyswatter” and the “lacrosse stick.” “The ground had faxed up to us some sketchy designs for these tools, and I think there were two tools that were made up. I can remember cutting up some plastic covers of some procedures books. We went around the cabin, all trying to find the piece parts, and the ground was helping us.”

Working together, the ground and the crew in space began an Apollo 13— like effort to improvise, using available materials to fashion a solution to the problem, Walker said. “The in-flight maintenance folks on the ground were, of course, very aware of what tools were on board, and they looked down the long list of everything that was manifest and tried to come up with a scheme of what pieces could be taken from here, there, and anywhere else on board, put together, and to make up these tools for swatting the satellite.”

The shuttle rendezvoused with the satellite, and Hoffman and Griggs exited the shuttle for the eva.

These guys go outside, and they’re oohing and aahing about the whole experience and doing great. . . . Rhea commands the remote manipulator system over to the side of the cargo bay. Literally with more duct tape and some cinching straps, they strap the flyswatter and the lacrosse stick on the end ofthe remote manipulator arm. Then they come back inside, and we make sure they’re okay, and they secure the suits. Bo and Don finish rendezvousing with the satellite, and Rhea very carefully moves the two tools on the end ofthe RMS right up against the edge ofthe satellite.

Walker noted that none of these procedures had been rehearsed on the ground; it was all improvised using the various skills the crew had picked up during their training. “This was all done just with the skills that the crew had been trained with generically, the generic operation of the remote manipulator system, the generic eva skills, and the generic piloting skills to rendezvous with another spacecraft,” Walker said. “And yet we pulled it off; the crew pulled it off expertly, did everything, including throwing the switch.”

Bobko said the crew had not done a rendezvous simulation or any ren­dezvous training in several months, and the books with rendezvous instruc­tions weren’t even on board. “So they sent us up this long teleprinted mes­sage, and I’ve got a picture of me at the teleprinter with just paper wound all around me floating there in orbit,” Bobko said. “It turned out to be a rather different mission. But, luckily, in training for the missions that had been scheduled before, we had learned all the skills that were required to do this. If we had just trained for this mission, we probably wouldn’t have ever trained to do a rendezvous or the other things that were required.”

Unfortunately, flipping the switch didn’t take care of the problem and there was nothing else the crew could do at that point. However, the ground was able to determine that the problem was with the electronics and the sat­ellite would need to be fixed on a subsequent flight. (It ultimately was re­paired on the 511 mission.) “We felt a little dismayed that the satellite failed on our watch and that we weren’t able to fix it on the same flight, but we felt gratified that we took one big step to finding out what the problem was, that eventually did lead to its successful deployment,” Walker said.

While the problems with the primary payload weren’t discovered un­til they got into the mission, another payload—Walker’s electrophoresis experiment—had encountered difficulties much earlier. About three days before launch, while the orbiter team was preparing Discovery for flight, Walker and several McDonnell Douglas folks were working with the cfes equipment when it started to leak. “My project folks were out there fill­ing it full of fluid, sterilizing it with a liquid sterilant, and then loading on board the sample material and then the several tens of liters of carrier fluid,” Walker said. “That electrophoresis device started leaking. Inside the orbiter, on the launchpad, it started leaking. Drip, drip, drip. Well, of course, that didn’t go over very well with anybody, and our folks diligently worked to resolve that. Right down to like twenty-four hours before flight or so, that thing was leaking out on the pad.”

Program managers began discussing whether the leak could be over­come so that the device could be loaded for operation. If not, the experi­ment could not be conducted during the mission. “The question became, ‘Well, maybe we don’t even fly Walker, if he doesn’t have a reason to fly,’” Walker recalled. “So there was active discussion until about a day before flight—this is all happening within about a twenty-four-hour period up till T minus twenty-four or thereabouts—as to whether I would fly or not, because maybe my device wasn’t going to be operational in flight and so I had nothing to do, so to speak. But it was resolved.”

The leak was fixed, the fluid was loaded, and the equipment—and Walker— were cleared for launch. In flight, the cfes worked well. And, in addition to running the cfes apparatus, during the mission Walker conducted the first protein crystal growth experiment in space, a major milestone in biotech­nology research. “This was the first flight of the U. S. protein crystal growth apparatus,” he said. “Actually, it was a small prototype that Dr. Charlie Bugg from the University of Alabama Birmingham and his then-associate, Larry DeLucas, had designed and had come to NASA, saying, ‘We’ve got this great idea for the rational design of proteins, but we need to crystallize these and bring the crystals back from space. We think they’ll crystallize much bet­ter in space, and we can do things up there we can’t do on Earth, etc., etc., but we need to fly it on board a Space Shuttle flight to see if it will work.’”

The flight was also the first for the NASA Education Toys in Space activ­ities, a study of the behavior of simple toys in a weightless environment. The project provided schoolchildren with a series of experiments they could do in their classrooms using a variety of toys that demonstrate the laws of physics. Astronauts conducted the experiments with the toys in orbit and videotaped their results. Students could then compare their results to what actually happened in space. The toys flown included gyroscopes, balls and jacks, yo-yos, paddle balls, Wheelos, and Hot Wheels cars and tracks. “I still to this day feel a little chagrined that I wasn’t offered a toy or the op­portunity,” Walker said. “Everybody else had a toy, but not me. . . . Even Jake Garn had paper airplanes.”

Walker may not have played with toys, but he played with liquids, con­ducting some fluid physics experiments with supplies on the orbiter.

Jeff Hoffman and I spent one hour preparing, at one point later in the mission, some drinking containers, one with strawberry drink and one with lemonade…. We would each squeeze out a sphere maybe about as big as a golf ball of liquid, floating in the cabin, and we actually played a little game in which we would put the spheres of liquid in free floating, oh, about a foot apart from each other, and Jeff and I would get on either side, and somebody would say, “Go." We’d start blowing at the spheres with our breath, just blowing on them, and we’d try to get them together and get them to merge, because it was really cool when they merged. One big sphere suddenly appears that’s half red and half green, and then the in­ternal fluid forces would start to mix them, and it’s really interesting to watch.

As the astronauts were blowing, their breath would actually move their bodies around. At the same time, the balls of liquid would start going in different directions, and the two together would make it increasingly dif­ficult to keep the liquid under control. “You’ve got to be quick,” Walker said, “and usually there’s got to be somebody with a towel standing by, be­cause either a wall or a floor or a person is going to end up probably get­ting some juice all over them.”

Walker said he felt more comfortable going in to his second flight than his first. “Not to say that I felt blase or ho-hum about it, by no means,” he said.

You just can’t go out and sit on a rocket and go into space and feel ho-hum about it, even after umpteen flights. It just isn’t going to happen. But a person can feel more comfortable. Some of the sharp edges, to put that term on it, of the un­known, of the tension, are just not there. I guess maybe a better way to put it, I would suggest, is now you really know when to be scared.

The second time around you’re not focusing on the same things. You’re now maybe a little less anticipatory of everything. You know [how] some things are going to be, so you can kind of sift those and put those aside in your mind and pay attention to other aspects. There were other things that I paid attention to, like I maybe was more observant of the Earth when I had a chance to look out the windows, more sensitive to the view.

Walker recalled Jeff Hoffman sharing with the crew his interest in as­tronomy, and in particular the crew trying to spot Halley’s Comet. “There was one or more nights, . . . in which we turned off all the lights in the cab­in and night-adapted our eyes. Everything was dark. . . . I can remember us trying to find the Halley’s Comet and never feeling like we succeeded at doing that. But, it was still so far away and so dim that it really probably wasn’t possible. But just looking at the sky along with an astronomer there was a great and tremendously interesting experience.”

Landing was delayed by a day, giving the crew an extra day in space, which Walker said he spent mostly looking out the window observing Earth. “I just never got bored at looking at the ever-changing world below,” Walker said. “You’re traveling over it at five miles per second, so you’re always see­ing a new or different part of the world, and even [as] days go by and you orbit over the same part of the world, the weather would be different, the lighting angles would be different over that part of the world. Just watch­ing the stars come up and set at the edge of the Earth through the atmo­sphere, watching thunderstorms.”

During the landing at Kennedy, Discovery blew a tire, resulting in exten­sive brake damage that prompted the landing of future flights at Edwards Air Force Base until the implementation of nose-wheel steering. Walker said the landing at first was just like the landing on his first mission. “Things were again just as they’d been before and as was planned and programmed, so no big surprises until those final few seconds when you expect to be thrown up against your straps by the end of the braking on the runway and the stop. Well, in our case, we’re rolling along about ready to stop, and then there’s a bang, and I can remember Rhea looking at me, and Jake saying, ‘What’s that?’”

Walker said one of the tires had locked up, skidded, and scuffed off a dozen layers of rubber and insulation and fiber until the tire pressure forced the tire to pop. “It ended up just a little bit off the center line of the run­way because of that, but we were going very slow, so there was no risk of running off the runway at that speed because of the tire blow. But certain­ly we heard it on board, and there was a thump, thump, thump, and we stop. We were going, like, ‘Well, what was that?’ I don’t know; in my own mind, I was thinking, ‘Did we run over an alligator? What happened here?’”

STS-51G

Crew: Commander Dan Brandenstein, Pilot John Creighton, Mission Specialists Shannon Lucid, John Fabian, and Steven Nagel, Payload Specialists Patrick Baudry (France) and Prince Sultan Salman Al-Saud (Saudi Arabia)

Orbiter: Discovery

Launched: 17 June 1985

Landed: 24 June 1985

Mission: Deployment of three communications satellites, test of spartan-i

Like so many missions before it, 51G succumbed to mission, crew, and payload shuffling. Commander Dan Brandenstein said shuffling like that was just how things were at this point in the shuttle program. There was a lot of scrambling around with missions for a variety of reasons, and the program was still relatively new, Brandenstein said.

That was early ’85. We had only been flying four years. The vehicle hadn’t ma­tured as you see it today. So they were flying technical problems on a vehicle and they’d have to pull one off the pad. That affected shuffling and payloads didn’t come along quite like they figured, and that affected shuffling. So it was sort of a variety of things. .. . Then we got canceled and picked up these four satellites. We had one for Mexico, one for the Arab Sat Consortium, one for AT&T, and then we had spartan, which was run out of Goddard. It was one that we de­ployed and then came back and recovered two days later. So it was a lot of mis­sion planning changed and we had a couple new crew people that we had to integrate into the crew and all that.

With three satellites to deploy into orbit, the 51G crew deployed one sat­ellite a day for the first three days on orbit. “Shannon and I had the lead on those deployments and J. O. Creighton was flying the orbiter, so he was pointing it in the right directions and so forth,” recalled Mission Specialist John Fabian. “Brandenstein was making sure that everybody was doing the right things. That’s what a commander is supposed to do. And Sultan was taking pictures for his satellite. I mean, it was a fairly routine operation.” The spartan proved to be a little more challenging. The spartan space­craft were a series of experiments carried up by the Space Shuttle. The pro­gram was based on the idea of a simple, low-cost platform that could be deployed from the Space Shuttle for a two – to three-day flight. The satel­lite would then be recovered and returned to Earth.

“It was a much simpler satellite,” Fabian said, “from the crew’s perspec­tive, than the SPAS-01 [a German satellite that Fabian released and recap­tured using the robotic arm on STS-7] because the SPAS-01, we could ma­neuver it. It had experiments on it that we could operate, had cameras on it that we could run. The spartan, which was a navy satellite, we simply released it, let it go about its business, and then later went back and got it.” Shannon Lucid did the release, and two days later Fabian did the recap­ture. Deployment was routine, Fabian said. “At least it appeared to be,” ex­plained Fabian. “When we left it, it was in the proper attitude. It was an x – ray astronomy telescope, and while we were gone, it took images of a black hole, which is kind of cool stuff. That’s kind of sexy.”

But when the orbiter came back to retrieve it, Fabian said, the satellite was out of attitude. The grapple fixture was in the wrong position for the shuttle’s arm to be able to easily grab it.

One idea was to fly an out-of-plane maneuver, flying the shuttle around the satellite, but the crew hadn’t practiced anything like that. Fabian noted,

Dan’s a very capable pilot, and I’m sure that he could have done that, but it turns out that perhaps an easier way would be to fly the satellite in much closer to the shuttle, get it essentially down almost into the cargo bay, and then reach over the top with the arm and grab it from the top, and that’s what we elected to do.

Of course, we told the ground what was going on, that it was out of atti­tude, and they worried, but there wasn’t much they could do—they couldn’t put it in attitude—so they concurred with the plan, and that’s what we executed.

Fabian said it felt good to benefit from all of the time spent in the simu­lators with the robotic arm. Training for contingency situations contribut­ed greatly to the crew’s knowledge of the arm’s capabilities and to the suc­cessful retrieval, he said.

The seven-person crew for this mission, including two payload specialists and representing three different nationalities, had a unique set of challeng­es because of those factors, said Fabian, but in general the crew got along well and had a positive experience.

We were told not to tell any camel jokes when Sultan showed up, and the first thing he did when he walked through the door was to say, “I left my camel out­side. ” So much for the public affairs part of the thing. These just were not issues. They really were not issues. Patrick flew a little bit ofFrench food and didn’t eat the same diet that we ate. Sultan did. Patrick flew some small bottles of wine that were never opened, but the press worried about whether or not they had been. Patrick flew as a Frenchman and enjoyed it, I think.

Fabian and Nagel were assigned to support Baudry and Sultan with any help they needed on their experiments. “Patrick was doing echocardio – graphs,” Fabian said, “and he did those on himself, and he did them on Sultan, and I think he did them on one or two of the nasa crew mem­bers, and frankly, I’ve forgotten whether he did one on me or not. But he was using a French instrument with a French protocol, and it was the principal thing that he was doing in flight, was to do these French medi­cal experiments.”

Sultan’s primary role was to observe, Fabian explained.

We were flying $130 million worth of satellites for the Arab League. But he also had some experiments, and he was tasked to take pictures, particularly over Sau­di Arabia, which of course would be very valuable when he got home. People would be very interested in seeing that. But they didn’t need a lot of support. They didn’t need a lot of help. We had to worry a bit about making sure that they were fed and making sure that they knew how to use the toilet and mak­ing sure that they understood the safety precautions that were there and so forth. And, you know, probably more than half of what our role and responsibility was with regard to the two. Other than that, it was to make sure they had film when they needed it in the cameras and help them for setup if they needed some setup for video or something of that type and to participate in their experiments to the degree that it was deemed necessary.

Even at this point, the payload specialist classification was still very new and crew members were still figuring out exactly how to act toward each oth­er, and that resulted in a change being made to the orbiter. “People weren’t really sure how these folks were going to react,” Fabian recalled. “We put a lock on the door of the side hatch. It was installed when we got into orbit so that the door could not be opened from the inside and commit hari-kari, kill the whole crew. That was not because of anybody we had on our flight but because of a concern about someone who had flown before.”

Fabian expressed concern over how the agency handled safety during this era of the shuttle program. On this flight, for example, Fabian said the arabsat never passed a safety review. “It failed every one of its safety re­views,” Fabian said.

The crew recommended that it not be flown, the flight controllers recommend­ed that it not be flown, and the safety office recommended that it not be flown, but NASA management decided to fly it. This was an unhealthy environment within the agency. We were taking risks that we shouldn’t have been taking. We were shoving people onto the crews late in the process so they were never fully integrated into the operation of the shuttle. And there was a mentality that we were simply filling another 747 with people and having it take off from Chica­go to Los Angeles, and this is not that kind of vehicle. But that’s the way it was being treated at that time.

It was very disappointing to a lot of people, a lot of people at the agency, to see management decide to fly this satellite. But if they hadn’t flown the satel­lite, you see, political embarrassment, what are we going to do with the Saudi prince, what about the French astronaut, what’s the French government going to have to say about us saying that we can’tfly their satellite on the shuttle, what will be the impact downstream of other commercial ventures that we want to do with the shuttle? Well, of course, after Challenger, the commercial all went away, and it was a dead-end street anyhow, but we didn’t know it at the time.

The Golden Age

How different it was in those early years of shuttle, when
we were going to fly once a month at least. That was going
to be routine, and we were going to revolutionize space and
discover these amazing things, and we still will, but we were
just naive, thinking it was going to happen the next year, and
not the next decade or the next generation. So there was a lot
of naivete, and maybe it was just us or maybe it was just me,
but that was the big change. It’s a little sad that that had to
happen, but that’s just maturing the industry, I guess.

—Astronaut Mike Lounge

STS-51I

Crew: Commander Joe Engle, Pilot Dick Covey, Mission Specialists Ox van Hoften, Mike Lounge, and Bill Fisher Orbiter: Discovery Launched: 27 August 1985 Landed: 3 September 1985

Mission: Deployment of three satellites; retrieval, repair, and redeploy­ment of syncom iv-3

Like the 51A mission, which recovered two satellites that had previously failed to deploy, 511 included the repair of a malfunctioning satellite from a previous mission, 51D. The satellite, syncom iv-3, had failed to activate properly after deployment. Mike Lounge recalled that he and fellow 511 mission specialist Ox van Hoften were together when they heard about the syncom failure. The sat­ellite was fine; the failure was a power switch on the computer. “We, essentially

on the back of an envelope, said, well, what’s the mass properties of this thing? Could it be handled by some sort of handling device by hand? Attached to the robot arm? And then if we had to push it away, what kind of forces would we have to push on it to make it stable, and is that a reasonable thing to do? So we calculated a twenty – or thirty-pound push would be enough.”

The calculations were right on. Lounge said computer simulations cal­culated a push of 27.36 pounds would be needed. Then the question was whether the rescue mission was even feasible. After looking into the chal­lenges further, Lounge and van Hoften believed it was and encouraged their commander to seek approval for his crew to do the job. Joe Engle shep­herded the request through center management and up to NASA headquar­ters and got the go-ahead for the recovery. Explained Lounge, “The key to the success of that mission and being able to do that was NASA was so busy flying shuttle missions that year that nobody was paying attention. If we’d had more attention, there’d have been a hundred people telling us why it wouldn’t work and it’s too much risk. But fortunately, there was a twelve­month period we flew ten missions; we were one of those.”

Discussions about the feasibility of the syncom recovery naturally led to comparisons with an earlier flight. One of the questions, recalled 511 pi­lot Dick Covey, was, would the astronauts be able to stop the rotation of the spacecraft? Covey said crewmate Ox van Hoften drew the solution on the back of a piece of paper.

He says, “It’s only going to take this much force to stop the rotation, so that’s not an issue. "Then [they] said, “Well, you know, does anybody think that we could have a person stop the rotation and do that?" Ox says, “Well, here. Here’s me," and he draws this big guy, and he says, “Here’s the syncom. " He draws a little guy, and he says, “Here’s Joe Allen, and there’s a palapa [satellite]. So if he can grab that one, then I can grab this one." We said, “Okay, yes." It was the “big astronaut, little astronaut" approach to things.

For Commander Engle, this mission would be very different from STS-2, on which he also served as commander. Engle described his second shuttle flight as less demanding than his first; the biggest difference between the two flights, he said, was that on his second mission—and NASA’s twentieth— there were more people there to help out.

We had only a crew of two on sts-2, and one of the lessons we learned from those first four orbital flight tests was that the shuttle—the orbiter itself—probably represents more of a workload than should be put onto a crew of two. It’s just too demanding as far as configuring all of the systems and switches, circuit break­ers. There are over fifteen hundred switches and circuit breakers that potentially have to be configured during flight, and some of those are in fairly time-critical times. . . . Some of them are on the mid-deck and some are on the flight deck, so you’re going back and forth and around. Having more people on board re­ally reduces the workload of actually flying the vehicle.

In flight, Engle discovered that having more crew to do the work meant more time personally to look out the window at Earth. “It’s a very, very in­spiring experience to see how thin, how delicate the atmosphere is, how beautiful the Earth is, really, what a beautiful piece of work it is, and to see the features go by,” Engle described.

Sultan Al-Saud was assigned to our crew initially, when one ofour payload sat­ellites was the arabsa T. [Al-Saud flew as part of the 51G crew.] He was assigned as a mission specialist on our crew, and when he eventually did fly, I think he said it better than anybody has. He said, “The first day or two in space, we were looking for our countries. Then the next day or two, we were looking at our continents. By about the fourth or fifth day, we were all looking at our world." Boy, it’s one of those things that I said, “God, I wish I’d have thought of that. I wish I’d have said that."

The crew made several launch attempts before finally getting off the ground on 27 August 1985. The first launch attempt, on 24 August, was scrubbed at T minus five minutes due to thunderstorms in the vicinity.

“When we got back in the crew quarters after that first scrub for weath­er,” recalled Covey,

[there was] John Young, who was the chief of the Astronaut Office at the time and also served as the airborne weather caller in the Shuttle Training Air­craft. . . . I was making some comments about, “I cant believe we scrubbed for those two little showers out there. Anybody with half a lick of sense would have said, ‘Let’s go. This could be a lot worse. ’"John Young came over and looked at me and he says, “The crew cannot make the call on the weather. They do not

know what’s going on. All they can see is out the window. That’s other people’s job. ” I said, “Yes, sir. Okay. ”

Launch was pushed to the next day, 25 August. Engle’s birthday was 26 Au­gust, and on board the shuttle was a cake that the crew was taking into orbit to celebrate. The launch scrubbed again on the twenty-fifth, this time because of a failure with one of the orbiter’s on-board computers, and was pushed to the twenty-seventh. Said Covey, “They wound up unstowing the birthday cake and taking it back to crew quarters, and we had it there instead of on orbit.”

Initially, things weren’t looking good for the third launch attempt, ei­ther, Engle said. It was raining so hard the crew wore big yellow rain slick­ers from the crew quarters to the Astrovan and up the elevator to the white room. Engle admitted that as the crew boarded, the astronauts “didn’t think there was a prayer” of actually flying that day. But they had only one more delay before they would have to detank and refuel the shuttle, which would delay the mission an additional two days, so the decision was made to get ready and see what happened.

We got in the bird and we strapped in and we started countdown. Ox van Hoften was in the number-four seat, over on the right-hand side aft, and Mike Lounge was in the center seat aft, and we were sitting there waiting, and launch con­trol had called several holds. Ox was so big that he hung out over the seats as he sat back, and he was very uncomfortable, and he talked Mike into unstrapping and going down to the mid-deck so that he could stretch across both those seats in the back of the flight deck. We were lying there waiting, and it was raining, and raining fairly good.

As the countdown clock ticked down nearer and nearer to the sched­uled launch time, the crew continued to believe that there was essentially no chance of a launch that day.

We got down to five minutes or six minutes, and. .. we got the call from launch control to start the apus. Dick Covey and I looked at each other kind ofincred – ulously and asked them to repeat. And they said, “Start the apus. We don’t have much time in the window here. ” So he started going through the procedures to start the apus, and they make kind of a whining noise as they come up to speed. The rest of the crew was asleep down in the mid-deck.

I think it was Fish [Bill Fisher] woke up and said, “What’s that noise? What’s going on?” We said, “We’re cranking apus. Let’s go,” or something like that. Dick was into the second apu, and they looked up and saw the rain coming down and they said, “Yeah, sure, we’re not going anywhere today. Why [are] you starting apus?” We didn’t have time to explain to them, because the sequence gets pretty rushed then. So we yelled to them, “Damn it, we’re going. Were going to launch. Get back in your seats and get strapped in. ” They woke up Ox and Mike, and they got back in their seats, and they had to strap themselves in. Normally you have a crew strap you in; they had to strap each other in. And Dick and I were busy getting systems up to speed and running, and all we could hear was Mike and Ox back there yelling at each other to, “Get that strap for me. Where’s my comm lead?” “Get it yourself. I cant find mine. ” And they were trying to strap them­selves in, and we were counting down to launch. They really didn’t believe we were going to launch because it was, in fact, raining, but they counted right down to the launch and we did go. It went right through a light rain, but it was raining.

Engle said that after the crew returned to Earth at the end of the mission he asked about the decision to launch through the rain. It turns out that the weather spotters were flying at the Shuttle Launch Facility at launch time, and it was clear there. “[They asked,] ‘Why didn’t you tell us it was raining [at the pad]?’ We used their rationale then. We said, ‘Our job is to be ready to fly. You guys tell us when the weather’s okay.’”

In addition to capturing and repairing the syncom that malfunctioned on 51D, the mission was to deploy three other satellites. One of those was another syncom, known as syncom iv-4 or leasat-4; the other two were Asc-i, launched for American Satellite Company, and the Australian com­munications satellite AussAT-i.

“We were supposed to do one the first day, one the second day, one the third day, and then the fourth and fifth days were repair days, and there was a day in between,” explained Lounge. But first, the crew was tasked with us­ing the camera to look at the payload bay and the sun shield to make sure everything was intact after launch. Lounge did just that.

Then I commanded the sun shield open, and I had failed to stow the camera. If it had been day two instead of day one, I’d have been more aware of it. On day one you’re just kind of overwhelmed and you’re just down doing the steps, and it’s not a good defense, but that was an example of why you don’t change things at the last minute and why you don’t do things you haven’t simulated, because we’d never simulated that. That was some engineer or program manager said, “Wouldn’t it be nice to add this camera task." Now I had a camera out of posi­tion, opened the sun shield against the camera, and it bent the sun shield and it got hung up on the top of the shuttle.

To address the problem, the crew did an earlier than planned checkout of the robotic arm and then used the arm to essentially bang against the sunshade, Lounge said. “[The sunshade was a] very flimsy structure with [an] aluminum tube frame and Mylar fabric, so not a lot to it, but it had to get out of the way.”

Lounge maneuvered the arm, but it wasn’t working right either.

The elbow joint had a problem that wouldn’t let the automatic control system operate the arm, so I had to command the arm single-joint mode, which means instead of some coordinated motion, command the tip to move in a certain tra­jectory, you just had to say, all right, elbow, move like this; wrist, move like this, rotate like this. So, a little awkward and took awhile, but I got the arm down there and banged on the solar array and got it down, and then we deployed that one [satellite]. . . . We deployed both of them on the same day, five or six hours after launch. So that was exciting, more exciting than it needed to be.

After the satellite deployments came the syncom repair attempt. The shuttle rendezvoused with the failed satellite, and robot arm operator Mike Lounge helped Bill Fisher and Ox van Hoften get ready for their eva. Once the eva began, van Hoften installed a foot platform on the end of the arm, and Lounge moved him toward the satellite so that he could grab it, just like in his napkin drawings months earlier.

The satellite, though, was in a tumble, so there were concerns about whether the capture could go as planned. As on earlier missions, the astro­nauts took advantage of the long periods of loss of signal and worked out a solution in real time.

“When we got up there and it was tumbling,” said Covey,

we were trying to relay back to the ground what was going on.. .. We were try­ing to figure out stuff. They were trying to figure out stuff. Finally we went LOS and Ox said, “Fly me up to it," and he went up and he just grabbed it. If the ground would have been watching, we wouldn’t have done that, I’m sure, like

that. But he grabs it and spins it, just with his hands on the edge, where they say, “Watch out for the sharp edges." And he spins it a little bit so that the fix­tures come around to him, and then he rotates it a little bit, and he gets that tool on, screws it down. We maneuver it down. We come aos. We say, “Well, we got it, Houston. "They didn’t ask why. They didn’t ask how.

STS-61B

Crew: Commander Brewster Shaw, Pilot Bryan O’Connor, Mission Spe­cialists Mary Cleave, Jerry Ross, and Sherwood Spring, Payload Spe­cialists Rodolfo Neri Vela (Mexico) and Charlie Walker

Orbiter: Atlantis

Launched: 26 November 1985

Landed: 3 December 1985

Mission: Deployment of three satellites, demonstration of space assem­bly techniques

As NASA worked to create a healthy manifest of shuttle flights, glitch­es with satellites’ inertial upper stages and with payloads themselves made for an ever-shifting manifest. “Continuously, we were juggling the mani­fest,” said astronaut Jerry Ross. “Crews were getting shifted from flight to flight. The payloads were getting shifted from flight to flight. And basical­ly, throughout 1985, our crew trained for every mission that flew that year except for military or Spacelab missions.”

At one point, Ross said, the 6ib crew was even assigned to 51L, the ill – fated final launch of the Space Shuttle Challenger. After bouncing around to several different mission possibilities, the crew settled in on 6ib.

The mission included two payload specialists, Charlie Walker and Ro­dolfo Neri Vela of Mexico. Ross remembered that the agency was being pressured to fly civilians—teachers, politicians, and the like. “We were giving away seats, is the way we kind of saw it, to nonprofessional astro­nauts, when we thought that the astronauts could do the jobs if properly trained,” Ross said.

The flight was Walker’s third flight in fifteen months and it was the first flight for a payload specialist from Mexico. “The guys did a great job on or­bit,” Ross praised the mission’s two payload specialists. “They were always very helpful. They knew that if the operations on the flight deck were very hectic, they stayed out of the way, which is the right thing to do, frankly.

But at other times they would come up onto the flight deck and enjoy the view as well as any of the rest of the crew.”

The mission deployed three more communications satellites: one for Mexico, one for Australia, and one for rca Americom. All three satellites were deployed using Payload Assist Modules. Ross and Mission Specialist Sherwood Spring worked together on the deployments. The two also jour­neyed outside the shuttle on two spacewalks to experiment further with as­sembling erectable structures in space. The two experiments were the Ex­perimental Assembly of Structures in Extravehicular Activity (ease) and the Assembly Concept for Construction of Erectable Space Structure (access).

“I’ll remember the day forever, when I got to go do my first spacewalk,” recalled Ross, who throughout his career ventured out of his spacecraft for a total of nine evas. That first venture outside was something he had been looking forward to for quite some time.

I got a chance to do a lot of[support for] spacewalks as a CapCom on the ground, and I got a little bit more green with envy every time I did that, thinking about what those guys were doing, how much fun they were having. So when I ulti­mately got a chance to go outside for my first time, I was worried, because I was worried that the orbiter was going to have a problem, we were going to have to go home early, or one of the spacesuits wouldn’t check out and we wouldn’t be able to go out, and all those things.

I’ll never forget opening up the hatch and poking my head out the first time, and I literally had this very strong desire to let out this war whoop of glee and excitement. But I figured that if I did that, they’d say, “Okay, Ross has finally lost it. Let’s get his butt back inside, ” and that would have been it. But it felt totally natural, just totally natural to be outside in your own little cocoon, your own little spacecraft, and I felt basically instantly at home in terms of going to work.

ease and access were designed to test how easily—or not—astronauts could assemble or deploy components in orbit. The idea was to study the feasibility of packing space structure truss components in a low-volume manner for transport to orbit so that they could then be expanded by as­tronauts in space. The question was more than just hypothetical; planning was already underway to construct a space station.

The idea of a space station was an old one, and nasa had already built and flown one space station, Skylab. When the Space Shuttle was proposed

The Golden Age

30. Jerry Ross on a spacewalk demonstrating the first construction of large structures in weightlessness. Courtesy nasa.

in the late 1960s, nasa’s desire was to build both the shuttle and a space sta­tion as the first steps in developing an infrastructure for interplanetary ex­ploration. The administration of President Richard Nixon approved fund­ing only for the Space Shuttle, however. By 1984, Reagan believed that the Space Shuttle program was sufficiently mature to move ahead with a space station, and the Space Station Freedom project was born. The station had ambitious goals, with plans calling for it to be a microgravity science lab, a repair shop for satellites, an assembly port for deep-space vehicles, and a commercial microgravity factory. Unlike Skylab, which was launched ful­ly assembled atop a powerful Saturn V rocket, plans were for Freedom to be launched in multiple modules aboard the shuttle and assembled in or­bit. The 61B spacewalks to test ease and access were part of the prepara­

tions, designed to study how best to build components so that they could be flown compactly and easily assembled by astronauts in space.

“The second spacewalk, we worked off the end of the mechanical arm for a lot of the work,” Ross explained.

We did the assembly, the top bay of the A ccess truss off the end of the arm. We simulated the running of the electrical cable. We did the simulation of doing a repair of the truss by taking out and reinserting an element there. We removed the trusses off of the fixture and maneuvered them around to see how that would be in terms of assembling a larger structure. We also mounted a U. S. flag that we had modified onto the truss and took some great pictures of us saluting the flag on the end of the arm up there. We also made a flag that we took outside. We called ourselves the Ace Construction Company. There’s a series of Ace signs that were taken outside on various spacewalks. . . . Somehow we’ve lost some of that fun over the years. I’m not sure why.

For the access work, the rms arm, with the spacewalkers at its end, was operated by Mary Cleave, whose height, or lack thereof, required spe­cial accommodation for her to be able to perform the task, according to Commander Shaw.

In order for her to get up and be able to look out the window and operate the controls on the rms, we’d strung a bungee across the panel and she’d stick her legs in front of that bungee and it would hold her against the panel . . . so she could be high enough and see and be in the right position to operate the rms. I remember coming up behind Mary once when she was operating the rms and there was somebody on the end of the arm. I put my hand on her shoulder, and her whole body was quivering, because she was so intent on doing this job right and not hurting anybody, and so focused and so conscientious, not wanting to do anything wrong, because she knew she had somebody out there on the end of this arm, and she was just quivering, and that just impressed the hell out of me, because I thought, you know, what a challenge, what a task for her to buy into doing when it obviously stressed her so.

All in all, Ross said, the two experiments were successful in their goal of producing data about in-space construction. “It gave us quite a bit of un­derstanding and knowledge of what it would be like to assemble things in space,” Ross noted.

Ultimately, that’s not the way that we chose to build the station, because when you think about having to integrate all the electrical and fluid lines and every­thing else into the structure once you’ve assembled this open network of truss, it becomes harder to figure out how you’re going to do that and properly con­nect everything together and make sure it’s tested and works properly. But we did learn a lot about assembling things in space and proved that they are val­id things that you could anticipate doing, even on the current station, at some point, if you needed to add a new antenna or something like that.

One advancement that came out of the access and ease experiments wasn’t even in space but had a big impact down the road. nasa realized through training for the space assembly tests that the Weightless Environ­mental Training Facility, or wet-f, water tank where astronauts were train­ing for eva was not going to be large enough to train for construction of a large space station. “[In] the facility we had when we built the access truss, we could only build like one and a half bays before it started stick­ing out of the surface of the water,” recalled Ross. “And the ease experi­ment, when we did it, basically our backpacks of our suits when we were at the top of the structure were right at the surface of the water. So if you’re going to build anything that’s anywhere close to being big on orbit, that wasn’t going to get it.”

For the next ten years, Ross helped nasa campaign to Congress for funds for a new facility. Ross helped design the requirements for the facility and led the Operational Readiness Inspection Team that eventually certified the new Neutral Buoyancy Laboratory at the Sonny Carter Training Facility.

During one of the two access/ease spacewalks, Ross recalled saying to Spring, “Let’s go build a space station.” Ross would later have the oppor­tunity to repeat that same phrase on his final spacewalk, on sTs-110 during actual assembly of the International Space Station in 2002.

The crew of 61B was the first to be on the shuttle in orbit on Thanks­giving Day, which meant, of course, that the astronauts needed a space – compatible Thanksgiving feast. Payload Specialist Charlie Walker recalled that the crew worked with jsc foods manager Rita Rapp on planning a spe­cial meal for the holiday. Rapp had been involved in space food develop­ment and astronaut menus since the early Mercury missions. Walker said the crew specifically requested pumpkin pie.

Of course, the menu had to be approved by NASA to withstand launch, and pumpkin pie didn’t make the cut. “Apparently somebody did the jiggle test, the vibration test, on the pumpkin pie, and what we were told later was, ‘Well, pumpkin pie does not make it to orbit. The center of the pumpkin pie turns back to liquid, so you won’t have pumpkin pie, you’ll have pumpkin slop in orbit, and you really don’t want that, so sorry, no pumpkin pie.’ So Rita said, All right, how about pumpkin bread? We can do that, and that will work, we know.’ So we had pumpkin bread on orbit for Thanksgiving.”

Thanksgiving was not the only interesting part of the mission from a culinary perspective. Being from Mexico, Payload Specialist Rodolfo Vela brought into space with him foods from home, one of which significantly changed the way astronauts would eat from that point forward. “Rodolfo had, of course, the desire, and probably the need, as it was perceived back home from Mexico, to be seen to be flying with some local Mexican cul­tural things, and so food was one of those,” Walker said.

One of the things that Rodolfo wanted to fly with, of course, was flour tortillas. In retrospect, I think that this amounted to something of a minor revolution in the U. S. manned space program, in that up to that time, of course, when crews wanted to have sandwiches in orbit, well, you went into the pantry, and you took out the sliced bread, sliced leavened bread that had been flown, for your sandwiches. Well, sliced bread, of course, always results in some degree of crumbs, and the crumbs don’t fall to the floor in the cabin in space. They are all around you, in your eyes, in your hair, and so it’s messy and just not that attractive. The crew saw Rodolfo flying with these flour tortillas and immedi­ately thought, “Ooh, this may be real good," and it was real good. It was tasty, after all, but when you took spread or anything that you wanted to make into a rolled sandwich and devoured it that way, but it was just no-muss, no-fuss kind of thing. I remember taking some sliced bread, but there may have been some sliced bread that even made it home, because we just found that the flour tortilla thing was well in advance of sliced bread, crumbly bread, for the prep­aration of sandwiches or just as a bread to go with your meal. The flour torti­llas worked well, much better than that.

Pilot Bryan O’Connor played a prank on Mission Specialist Spring. Spring was in the army—a West Point grad—and O’Connor was from the

Naval Academy. During the mission the two armed forces faced off in the annual Army-Navy Game. O’Connor’s prank centered around the rivalry between the two forces.

Each person was allowed to carry six audios, and NASA would help you record records or whatever you wanted onto these space-qualified audiotapes. Then we would carry them and a tape player on board with our equipment. Usu­ally what would happen is people would break those out when it was time for bed and listen to their favorite music at bedtime. . . . It was on day three that we turned off the lights and, I don’t know, it was about ten minutes after the lights were off, and I was borderline asleep, and I hear this loud cry from the other side of the mid-deck, where Woody [Spring] was hanging off the wall in his bed. He yells out, “O’Connor, you S. O.B.!" It woke me up with a start, and I had no idea what he was talking about. “What is it? What is it?" He says, “You know what it is."

And all of a sudden, it clicked with me. About a month before flight, when we were having the people transcribe music onto our tapes, I went over to the guy that was working on Woody’s tapes and I gave him a record with the Na­val Academy fight song on it and I said, “I want you to go right in the middle of his tape somewhere, just right in the middle, and superimpose the navy fight song somewhere on his tape." Well, it turned out it was his Peter, Paul, and Mary album, and it was right in the middle of “I’ve Got a Hammer." He’s lis­tening to “I’ve Got a Hammer" on his way to sleep and suddenly up comes this really loud navy fight song thing right in the middle of it. We still joke about that to this day. In fact, sometimes we go to one or the other house and watch the Army-Navy Game together, and we always remember that night on the At­lantis in the mid-deck.

The mission ended and the crew came in for landing at Edwards Air Force Base in California. Weather brought the crew in one orbit sooner than was originally planned. “We came to wheels stop, and everybody unbuckles, and they’re trying to get their land legs again,” said Walker.

Jerry [Ross] is over at the hatch real quickly and wants to pop the hatch open so that we’ve got that part of the job done. Well, Jerry pops the hatch open, but it literally pops open, because whomever had planned these things had forgotten about the altitude, pressure altitude difference, between sea level at the Cape and the probably three-thousand-, four-thousand-foot elevation at Edwards Air Force Base. So it’s a little bit less air pressure outside. Well, we’re still at sea-level pressure inside the ship. So he turns the crank on the side hatch, and the hatch goes, “Pow!" It flops down, and right away, I think Jerry said something about, “Oh, my God, I’m going to have to pay for a new hatch."

STS-61C

Crew: Commander Hoot Gibson, Pilot Charles Bolden, Mission Special­ists Franklin Chang-Diaz, Steven Hawley, and Pinky Nelson, Payload Specialists Robert Cenker and Congressman Bill Nelson

Orbiter: Columbia

Launched: 12 January 1986

Landed: 18 January 1986

Mission: Deployment of the rca satellite satcom ku-i, various oth­er experiments

The payload specialists on 61c were Robert Cenker of rca, who dur­ing the mission observed the deployment of the rca satellite, performed a variety of physiological tests, and operated an infrared imaging camera; and the second member of the U. S. Congress to fly, Bill Nelson, a member of the House of Representatives, representing Florida and its Space Coast, and chairman of the Space Subcommittee of the Science, Space, and Tech­nology Committee.

Pinky Nelson recalled that the payload specialists assigned to the mis­sion changed several times leading up to flight. “Our original payload spe­cialists were Bob Cenker and Greg Jarvis, so they were training with us,” Pinky said.

It was after Jake Garn flew, and then they decided they had to offer a flight to his counterpart in the House. Don Fuqua couldn’t fly for some reason, and so it filtered down to the chair of the subcommittee, Bill Nelson, and he jumped at the chance. Who could blame him? This was just months before the flight, in the fall or late fall, even. The flight was scheduled in December. So they bumped Greg and his little payload off the mission over onto Dick Scobee’s [51L Chal­lenge^ crew and added Bill to our crew. I think our attitude generally at that point was, “Well, that’s just the way the program’s going. Were flying payload specialists. We’ll make the best of it."

Pinky Nelson described Representative Nelson as a model payload spe­cialist, working very hard to contribute to the mission. “He had no experi­ence either in aviation or anything technical. He was a lawyer, so he had a huge learning curve, but that didn’t stop him from trying, and I think he knew where his limitations were,” Pinky said. “He wanted to jump in and help a lot of times, but just didn’t have the wherewithal to do it, but worked very hard and was incredibly enthusiastic.”

The launch of 61c was delayed seven times. Originally set for 18 Decem­ber 1985, the launch was delayed one day when additional time was need­ed to close out the orbiter’s aft compartment. On 19 December the launch scrubbed at T minus fourteen seconds due to a problem with the right sol­id rocket booster hydraulic power unit. “As it turned out,” Charlie Bolden said, “when we finally got out of the vehicle and they detanked and went in, they determined that there wasn’t really a problem. . . . It was a com­puter problem, not a physical problem with the hydraulic power unit at all, and it probably would have functioned perfectly normally, and we’d have had a great flight.”

The launch was pushed out eighteen days, to 6 January 1986. The third attempt stopped at T minus thirty-one seconds due to the accidental drain­ing of approximately four thousand pounds of liquid oxygen from the ex­ternal tank. The next day launch scrubbed at T minus nine minutes due to bad weather at both transoceanic abort landing sites.

Two days later, on 9 January, launch was delayed yet again, this time be­cause a launchpad liquid oxygen sensor broke off and lodged in the num­ber two main engine prevalve. “That time we got down to thirty-one sec­onds, and one more time things weren’t right,” Bolden said. “So we got out, and it was another main engine valve. This time they found it. There had actually been a probe, a temperature probe, that in the defueling, they had broken the temperature probe off, and it had lodged inside the valve, keeping the valve from closing fully. So that would have been a bad day. That would have been a catastrophic day, because the engine would have exploded had we launched.”

A 10 January launch was delayed two days due to heavy rains. After so many delays, despite the adverse weather conditions, the crew was still load­ed onto the vehicle, on the off chance that things would happen to clear up. “It was the worst thunderstorm I’d ever been in,” Bolden said.

We were really not happy about being there, because you could hear the light­ning. You could hear stuffcrackling in the headset. You know, you’re sitting out there on the top of two million pounds of liquid hydrogen and liquid oxygen and two solid rocket boosters, and they told you about this umbrella that’s over the pad, that keeps lightning from getting down there, but we had seen light­ning actually hit the lightning-arrester system on sts-8, which was right there on the launchpad. So none of us were enamored with being out there, and we started talking about the fact that we really ought not be out here.

While some astronauts found multiple scrubs and launch attempts frus­trating, Steve Hawley said he wasn’t bothered by it. “My approach to that has always been, hey, you know, I’d go out to the launchpad every time ex­pecting not to launch,” Hawley said.

If you think about all the things that have to work, including the weather at several different locations around the world, in order to make a launch hap­pen, you would probably conclude, based on the numbers, that it’s not even worth trying. So I always figured that we’re going to turn around and come back. So I’m always surprised when we launch. So my mindset was always, you know, we’ll go out there and try and see what happens. So I never real­ly viewed it as a disappointment or anything. I always feel a little bad for, you know, maybe family and guests that may have come out to watch, that now they have to deal with the fact there’s a delay and whether they can stay, whether they have to leave, and that’s kind of a hassle for them, but it never bothered me particularly.

In those days, Hawley pointed out, the launch windows were much lon­ger than they were later in the program for International Space Station dock­ing missions. In the early days launch windows were two and a half hours, versus five to ten minutes for iss flights. The longer windows were advan­tageous in terms of probability of launch but could be adverse in terms of crew comfort. “You’re out there on your back for five or six hours, and that gets to be pretty long, day after day, but the fact that you didn’t launch nev­er bothered me particularly.”

Having been part of multiple scrubs on his earlier 41D mission, Haw­ley had a reputation for not being able to launch. “I don’t remember how we came up with the specifics of the disguise, but I decided that if it didn’t

know it was me, then maybe we’d launch, and so I taped over my name tag with gray tape and had the glasses-nose-mustache disguise and wore that into the [white] room. I had the commander’s permission, but I don’t re­member if we had told anybody else we were going to do that. . . . Evident­ly it worked, because we did launch that day.”

So finally, on 12 January 1986 the mission made it off the ground. By this time the crew members were wondering if they’d ever really go. “We did all kinds of crazy stuff,” Bolden said, “fully expecting that we wouldn’t launch, because I think the weatherman had given us a less than 50 percent chance that the winds were going to be good or something, so we went out and we were about as loose as you could be that morning. And they went through the countdown, came out of the holds and nothing happened. ‘Ten, nine, eight, seven, six.’ And we looked at each other and went, ‘Holy—we’re re­ally going to go. We’d better get ready.’ And the vehicle started shaking and stuff, and we were gone.”

Within seconds of lifting off, an alarm sounded. “I looked down at what I could see, with everything shaking and vibrating, and we had an indica­tion that we had a helium leak in—I think it was the right-hand main en­gine,” Pilot Bolden recalled. “Had it been true, it was going to be a bad day. . . . We didn’t have a real problem. We had a problem, but it was an instrumentation problem.”

With the determination made that the indicator was giving a false read­ing, the crew continued on into space. “We got on orbit and it was awe­some,” said Bolden of his first spaceflight. “It was unlike anything I’d ex­pected. Technically, we were fully qualified, fully ready, and everything. Emotionally, I wasn’t even close. I started crying. Not bawling or anything, but just kind of tears rolling down my cheek when I looked out the win­dow and saw the continent of Africa coming up. It looked like a big island. Just awesome, unlike anything I’d ever imagined.”

Pinky Nelson said he found the entire mission rather frustrating, de­scribing the mission as trivial. “We launched one satellite, and we did this silly material science experiment out in the payload bay which didn’t work. I knew it wasn’t going to work when we launched. Halley’s Com­et was up at the time and we had this little astronomy thing to look at Halley’s Comet, and it was launched broken, so it never worked. So the mission itself, to say what we did, I don’t know. I deployed a satellite.

Steve deployed a satellite. We threw a bunch of switches, took a bunch of pictures.”

According to Bolden, the crew referred to its mission as an end-of-year – clearance flight. “We had picked up just tons of payload, science payloads that Marshall [Space Flight Center, in Huntsville, Alabama] had been trying to fly for years, and some of the Spacelab experiments and stuff that they couldn’t get flown,” Bolden said. “So we had Congressman Nelson and ev­ery experiment known to man that they couldn’t get in. There was nothing spectacular about our mission. It was almost like a year-end clearance sale.”

Hawley said there was a general feeling at the time that without Nelson on the flight the mission might have been canceled. “We all suspected, al­though no one ever said, that because of the delays that we got into and the fact that, frankly, our payload wasn’t very robust, that were it not for his presence on the flight, we might have been canceled. . . . We wondered about that and always thought, without knowing for sure, that that might have happened if we hadn’t had a congressman, but this was his flight, and so we had some guarantee that it would happen.”

In addition to launching the rca satellite satcom ku-i, the mission was assigned the Comet Halley Active Monitoring Program (champ) ex­periment, which was a 35 mm camera that was to be used to photograph Halley’s Comet. But the camera didn’t work properly due to battery prob­lems. Science experiments on board were the Materials Science Laborato­ry-2, Hitchhiker g-i, Infrared Imaging experiment, Initial Blood Storage experiment, Handheld Protein Crystal Growth experiment, three Shuttle Student Involvement Program experiments, and thirteen Getaway Specials.

“The big challenge,” Bolden said,

was arguing with the ground about how we should do some of the experiments. There were some that we could see were not going exactly right. I didn’t have the problem as much as Pinky. Pinky was the big person working a lot of the material sciences experiments. And while we had very little insight into what was going on inside the box, we could tell that because of the data that we were seeing on board, we could tell that if we were just given an opportunity to reenergize an experiment, or to turn the orbiter a different way, or do some­thing, we might be able to get some more data for the principal investigators. The principal investigators agreed, but the flight control crew on the ground [said] that wasn’t in the plan. They weren’t interested in ad-libbing. They had a flight to fly and a plan to fly, and so forget about these doggone experiments.

Bolden said there was generally always a power struggle between the flight director on the ground and the commander of the shuttle in space, an issue that dated back to the earliest U. S. spaceflights.

There’s always a pull and tug between the flight director, who is in charge— nobody argues that point—and the crew commander, who, by the General Pru­dential Rule ofSeamanship of navigation at sea, has ultimate responsibility for the safety of the crew and vessel. If there’s a disagreement between the command­er and the flight director—and that’s happened on very, very few occasions, but every once in a while it happens—the commander can do what he or she thinks is the right thing to do and is justified in doing that by the General Prudential Rule. And even NASA recognizes that. Now, you could be in deep yogurt when you come back, ifsomethinggoes wrong. But you have the right to countermand the direction of the flight director. Almost never happens.

Bolden recalled that because the agency wanted to get Challenger and the Teacher in Space mission off the ground, 61c was cut from six days to four. The flight was scheduled to land at Kennedy Space Center—which would have made it the first flight to land there since the blown-tire incident on 51D—but the weather in Florida once again didn’t cooperate.

“They kept waving us off and making us wait another day to try to get back into Ksc,” Hawley recalled. “What I remember is that by the third day we had sort of run out of most everything, including film, and part of our training had been to look for spiral eddies near the equator, because the theory was, for whatever reason, you didn’t see them near the equa­tor, and Charlie was looking out the window and claimed to see one, and I told him, ‘Well, you’d better draw a picture of it, because we don’t have any film.’ So we couldn’t take a photograph.”

After two days of bad weather preventing the shuttle from landing in Florida, it was decided to send Columbia to Edwards Air Force Base. “The first attempt at Edwards was waved off because the weather there was bad,” Bolden said. “Finally, on our fifth attempted landing. . . in the middle of the night on 18 January, we landed at Edwards Air Force Base, which was inter­esting because with a daytime scheduled landing, you would have thought that we wouldn’t have been ready for that. And Hoot, in his infinite wis­dom, had decided that half of our landing training was going to be night­time, because you needed to be prepared for anything. And so we were as ready for a night landing as we could have been for anything.”

The landing was smooth, but Bolden said Congressman Nelson was dis­appointed not to have landed in his home state. “He really had these visions of landing in Florida and taking a Florida orange or something, and boy, the crew that picked us up was unmerciful, because they came out with a big—it wasn’t a bushel basket, it was a peck—basket of California orang­es and grapefruits. And even having come from space, he was just not in a good mood. So that was a joke that he really did not appreciate.”

Pinky Nelson recalled that Bill Nelson was struggling with not feeling well after landing but kept going anyway. “Most people don’t feel very good their first day or two in space, but don’t have too much trouble when they get back on the ground,” Pinky said. “Bill had a really hard time for a few hours after we landed, but boy, he was a trooper. He was suffering, but he— you know, good politician—put on a good face, and we had to do our lit­tle thing out at Edwards and all that and get back on the plane. He really sucked it up and hung in there, even though he was barely standing. And the rest of us, of course, cut him no slack at all.”

Bolden described being in awe of Gibson’s skills as a pilot and told how the five-time shuttle veteran took Bolden under his wing. “The way that I was trained with Hoot was you don’t ever wing anything,” Bolden said.

I credit him with my technique as a commander. He preached from day one, “We don’t ever do anything from memory. We don’t ever wing it. If something’s going to happen, there is a procedure for it. And if there’s not a procedure for it, then we’re going to ask somebody, because somebody should have thought about it." And so what we did was we trained ourselves just to know where to go in the book. And hopefully, crews still train like that, although I always flew with people who would invariably want to wing it, because they prided themselves in having photographic memories or stuff like that. The orbiter and just space­flight is too critical to rely on memory, when you’ve got all of these procedures that you can use, and the ground to talk to.

Gibson taught him what Bolden referred to as Hoot’s Law, a piece of wisdom that has stayed with him ever since. “We were in the simulator one day,” Bolden said, “in the sms [Shuttle Motion Simulator], and I was still struggling. It was in my struggling phase. And I really wanted to impress everybody on the crew and the training team. We had an engine go out, boom, like that, right on liftoff.”

During training, Bolden explained, the sim crew would introduce some errors to see how the astronauts would respond to them, but then at other times, they introduced abnormalities simply to try to distract the astronauts from the important things. “There is probably one criti­cal thing that you really need to focus on, and the rest of it doesn’t make any difference. If you don’t work on it, you get to orbit and you don’t even know it was there. But if you notice it and start thinking about it or start working on it, you can get yourself in all kinds of trouble. They love doing that with electrical systems, so they would give you an elec­trical failure of some type.”

Bolden recalled that on this particular occasion he was working an en­gine issue, and the sim team introduced a minor electrical problem. He first made the mistake of trying to work the corrective procedure on the wrong electrical bus, one of multiple duplicate systems on the shuttle.

The training team intended it this way. You learn a lesson from it. So I started working this procedure and what I did in safeing the bus was I shut down the bus for an operating engine. When I did that, the engine lost power and it got real qui­et. So we went from having one engine down in the orbiter, which we could have gotten out of to having two engines down, and we were in the water, dead. Here I went from I was going to feel real good about myselfbecause I’d impress my crew to feeling just horrible because I had killed us all. And Hoot kind ofreached over and patted me on the shoulder. He said, "Charles, let me tell you about Hoot’s Law. ” That’s the way he used to do stuff sometimes. And I said, “What’s Hoot’s Law?” And he said, “No matter how bad things get, you can always make them worse. ”

And I remembered Hoot’s Law from that day. That was probably 1984, or 1985 at the latest, early in my training. But I remembered Hoot’s Law every day. I have remembered Hoot’s Law every day of my life since then. And I’ve had some bad things go wrong with me in airplanes and other places, but Hoot’s Law has always caused me to take a deep breath and wait and think about it and then make sure that somebody else sees the same thing I did. And that’s the way I trained my crews, but that was because of that experience I had with Hoot.

STS-62A

Crew: Commander Bob Crippen, Pilot Guy Gardner, Mission Special­ists Dale Gardner, Jerry Ross, and Mike Mullane, Payload Specialists Brett Watterson and Pete Aldridge

Orbiter: Discovery

Launched: N/A

Landed: N/A

Mission: Deployment of reconnaissance satellite, first launch from Van – denberg

As 1985 wrapped up and 1986 began, the Space Shuttle was beginning to realize its promised potential. The early flights had hinted at that prom­ise, but in 1985, with the orbiter’s high flight rate, variety of payloads, and distinguished payloads, the nation was beginning to see that potential be­come reality. And as those flights were taking place, work on the ground was foreshadowing even greater times ahead—an even greater flight rate, a teacher and journalist flying into space, the imminent launch of the Hub­ble Space Telescope, and entirely new classes of missions in planning that would mark even broader utilization for America’s versatile Space Trans­portation System.

After his completion of mission 41G, Bob Crippen sought and was giv­en command of a groundbreaking mission, 62A. The mission was the first time that the “2” was used in a designation, indicating that the launch was to take place not out of Kennedy Space Center but rather out of Vanden – berg Air Force Base in California.

“The air force built the launch site out there to do military missions which required a polar orbit, and it was a flight I wanted a lot,” said Crippen, who explained that the assignment was a homecoming of sorts for him. During the 1960s, before his transfer to NASA, Crippen had been a part of the air force astronaut corps, based at Vandenberg and assigned to the Manned Orbiting Laboratory space station program. The mol launches would have used Space Launch Complex 6 (slc-6) at Vandenberg, and now SLC-6 had been modified to support Space Shuttle launches.

“I felt like I’d come full circle, and I really wanted that polar flight,” Crip­pen said. “I lobbied for it and ended up being selected, although not with­out some consternation. I think since this was primarily an air force mis­

sion, there was a big push by the air force to have an air force commander on the flight. But the powers that be ended up discussing it a lot and let­ting me take the lead on it.”

The crew, which included U. S. undersecretary of the air force Pete Aldridge, spent a lot of time at Vandenberg, Crippen said, making sure the launch com­plex was acceptable. “We actually took the Enterprise out there and used it to run through where they had to move it to stack it, and they actually had an external tank and some not-real solid rockets out there. . . . So we mounted it all up, and I’ve got pictures of the vehicle sitting on the launchpad like it’s ready to fly, but it was the Enterprise, as opposed to the Discovery.”

Jerry Ross recalled that a key difference between Kennedy and Vanden­berg was that the shuttle stack was going to be assembled on the launchpad itself. At Kennedy, the orbiter, srbs, and external tank were stacked inside the massive Vehicle Assembly Building and then moved to the launchpad on the crawler. “The solid rocket motors were going to be stacked up out on the pad. The external tank would be mounted to those out at the pad, and then the shuttle [orbiter] would be brought out on this multitiered carrier from its processing facility several miles away and taken out to the launch – pad and put in place once everything else was ready,” said Ross. “The en­tire launch stack could be enclosed in basically a rollaway hangar type of facility, and also the Launch Control Center was basically underneath the pad. It was buried in the concrete, not directly underneath, but still right there contiguous to the launchpad itself. That should have been a fairly noisy place to operate out of.”

Ross learned that he had been assigned to 62A while still in training for 6ib. “I was assigned to a second flight before I flew the first one,” he said.

I was very excited about that, and the fact that you’re going to get to do some­thing so unique like that for the very first time was fascinating to me. When I launched in November of’85, I was supposed to fly again in January of’86, out of Vandenberg. Of course, everybody knew that date was not realistic at that point. But while I was on orbit, that date had been slipped out to July of ’86, and most people thought that that was a fairly realistic date. So that would have been very close, two flights within six, seven months of each other.

Training for 6ib kept him quite busy, but Ross managed to squeeze out a little extra time to also train with the 62A crew at Vandenberg. “I was wor­

ried with the flights getting so close together that maybe they were going to replace me,” Ross admitted. “I talked to Crip a couple of times about that, and he said, ‘Don’t worry. We’ll take care of you.’”

The planned launch, Ross said, would have been a fascinating ride. “We were going to go into a 72И degree inclination orbit. . . . It would have been awesome. We’d have basically seen all the land masses of the world, so it would have been neat.”

Fellow 62A crew member Mike Mullane agreed. “The idea of flying into polar orbit, oh, man, I was just looking forward to that so much. You’re ba­sically going to see the whole world. In an equatorial orbit like we were fly­ing, or a low-inclination orbit like we were flying on the first mission, you don’t get to see lot of the world. So I was really looking forward to that.” As with other DoD missions the shuttle had flown, preparations for 62A involved a high amount of secrecy, with astronauts required to not even tell their wives what they were doing. However, Mullane said he enjoyed working on military missions; his next two shuttle missions were also mili­tary missions. “You had a sense of this national security involved about it, which made you feel a little bit more pride, I guess, in what you were do­ing and importance in what you were doing.”

Ross said the original flight plan had included twenty-four-hour-a-day operations by two air force payload specialists—Brett Watterson and Ran­dy Odle—but Odle was bumped in favor of Pete Aldridge, undersecretary of the air force. “That would have been some pretty high-power folks fly­ing with us on the flight.”

The mission was assigned two main payloads: an experimental infrared telescope called Cirris and a prototype satellite, p-8888, called Teal Ruby. “My understanding was [Teal Ruby] was a staring mosaic infrared sensor satellite that was trying to be able to detect low-flying air-breathing vehicles, things like cruise missiles, and a way to try to detect those approaching U. S. territories.” Mullane said he had no additional concern at the prospect of launching from a previously unused launch complex. “Not any more beyond a natu­ral terror of riding a rocket,” said Mullane.

I don’t care where it was launching from; I didn’t personally have any fear about it being a new launchpad and therefore more danger associated with it. It’s just that on launch on a shuttle, you fly with no escape system: no ejection seat, no pod, no parachute of any form. You fly in a rocket that has a flight-destruct sys­tem aboard it, so it can be blown up in case something goes wrong. Those are reasons why you’re terrified. It’s not where you’re launchingfrom, in my opinion; it’s [that] the inherent act of flying one of these rockets is dangerous.

STS-61F

Crew: Commander Rick Hauck, Pilot Roy Bridges, Mission Specialists Mike Lounge and David Hilmers

Orbiter: Challenger

Launched: N/A

Landed: N/A

Mission: Launch of the Ulysses space probe

While Crippen and his crew were getting ready for 62A, Rick Hauck was preparing for his own first-of-its-kind Space Shuttle mission. He was the astronaut project officer for the Centaur cryogenically fueled upper stage, which nasa was planning to use on the shuttle as a platform for launch­ing satellites.

The Centaur upper stage rocket had a thin aluminum skin and was pres­sure stabilized, such that if it wasn’t pressurized, it would collapse under its own weight, like the Atlas missiles used to launch the orbital Mercury missions. “If it were not pressurized but suspended and you pushed on it with your finger, the tank walls would give and you’d see that you’re flex­ing the metal,” said Hauck. “Its advantage was that it carried liquid oxy­gen and liquid hydrogen, which, pound for pound, give better propulsion than a solid rocket motor [like NASA had been using on previous missions to boost satellites].”

Preparations were being made for the Centaur to launch two interplan­etary probes—the Ulysses probe and the Galileo probe—which needed the powerful rockets available to be launched into deep space from the shuttle.

“At some point,” Hauck said, “the decision was made, well, we’ve got to use the Centaur, which was never meant to be involved in human space­flight.” The origins of Centaur are older than NASA itself; it began as a project of the air force in 1957. Throughout its history it has been useful as an upper stage on expendable launch vehicles for launching satellites and probes to the moon and to planets other than Earth. But there was a high level of danger involved in pairing the highly volatile Centaur with a shuttle full of people.

“Rockets that are associated with human spaceflight have certain levels of redundancy and certain design specifications that are supposed to make them more reliable,” Hauck commented.

Clearly, Centaur did not come from that heritage, so, number one, was that go­ing to be an issue in itself, but number two is, if you’ve got a return-to-launch – site abort or a transatlantic abort and you’ve got to land, and you’ve got a rock­et filled with liquid oxygen/liquid hydrogen in the cargo bay, you’ve got to get rid of the liquid oxygen and liquid hydrogen, so that means you’ve got to dump it while you’re flying through this contingency abort. And to make sure that it can dump safely, you need to have redundant parallel dump valves, helium sys­tems that control the dump valves, software that makes sure that contingen­cies can be taken care of And then when you land, here you’re sitting with the shuttle Centaur in the cargo bay that you haven’t been able to dump all of it, so you’re venting gaseous hydrogen out this side, gaseous oxygen out that side, and this is just not a good idea.

Hauck was working those issues when George Abbey called on him to command the first flight of the Shuttle-Centaur to launch the Ulysses solar probe. Astronaut Dave Walker was assigned the second Centaur mission, to launch the Galileo probe to Jupiter. The two missions had to be flown close together—in the first ten days of April 1986, Hauck said—because of the positioning of Earth in its orbit relative to the two satellites’ destinations. “It was clear this would be very difficult,” Hauck said. “We were going to have just four crew members, because that minimized the weight. We were going to 105 nautical mile altitude, which was lower than any shuttle had ever gone to, because you need the performance to get the Shuttle-Centaur up because it was so heavy.”

The Shuttle-Centaur integration was being managed out of the Lewis Research Center, now Glenn Research Center, in Cleveland, Ohio, where the Centaur was developed originally. “Lewis had been the program man­agers for Atlas-Centaur, and so they knew the systems,” Hauck said,

but in retrospect, the whole concept of taking something that was never designed to be part of the human spaceflight mission, that had this many potential fail­ure modes, was not a good idea, because you’re always saying, “Well, I don’t want to solve the problems too exhaustively; I’d like to solve them just enough so that I’ve solved them." Well, what does that mean? You don’t want to spend any more money than you have to, to solve the problem, so you’re always try­ing to figure out, “Am I compromising too much or not?" And the net result is you’re always compromising.

The head of the Office of Spaceflight at that time was Jess Moore, whom Hauck described as a good man but one who was unfamiliar with the world of human spaceflight. “Jess made it very clear that he wanted Dave and my­self to be part of all the substantive discussions, and he was very sensitive to the human spaceflight issues, but he wasn’t a human spaceflight guy,” Hauck said. “I think that the program would have profited at that point by having had someone there who was more keenly attuned to the human spaceflight issues. As I say, he couldn’t have been nicer to us and encour­aged us more and bent over backwards to be sensitive to the issues, but he didn’t start out as a human spaceflight guy.”

In early January 1986, Hauck recalled, he worked on an issue with re­dundancy in the helium actuation system for the liquid oxygen and liquid hydrogen dump valves. It was clear, in Hauck’s mind, that the program was willing to compromise on the margins in the propulsive force being provid­ed by the pressurized helium, which concerned him enough that he took it up with Chief of the Astronaut Office John Young. “John Young called this mission ‘Death Star,’” recalled Hauck. “That was his name for this mission, which he said with humor, but behind humor, there’s a little bit of truth. I think it was conceded this was going to be the riskiest mission the shuttle would have flown up to that point.”

Young, Hauck, and other members of the Astronaut Office argued be­fore a NASA board why it was not a good idea to compromise on this fea­ture, and the board turned down the request. “I went back to the crew of­fice and I said to my crew, in essence, ‘nasa is doing business differently from the way it has in the past. Safety is being compromised, and if any of you want to take yourself off this flight, I will support you.’”

Hauck said he didn’t consider asking to be removed from the mission himself.

I probably had an ego tied up with it so much that, you know, “I can do this. Heck, I’ve flown off of aircraft carriers, and I’ve flown in combat, and I’ve put myself at risk in more ways than this, and I’m willing to do it. ” So I didn’t ever think of saying, "Well, I’m not going to fly this mission. ” Knowing what I know now, with Challenger and Columbia, maybe I would. But NASA was a lot different back there, when we’d never killed anybody in spaceflight up to that point. I mean, there was a certain amount of sense that it wouldn’t happen.

To Touch the Face of God

And, while with silent lifting mind I’ve trod
The high untrespassed sanctity of space,

Put out my hand and touched the face of God.

—“High Flight,” Pilot Officer Gillespie Magee, No. 412
Squadron, Royal Canadian Air Force,
died 11 December 1941

sts-5il

Crew: Commander Dick Scobee, Pilot Michael Smith, Mission Special­ists Ellison Onizuka, Judy Resnik, and Ron McNair; Payload Special­ists Christa McAuliffe and Gregory Jarvis Orbiter: Challenger Launched: 28 January 1986 Landed: N/A

Mission: Deployment of TDRS, astronomy research, Teacher in Space

Astronaut Dick Covey was the ascent CapCom for the 51L mission of the Space Shuttle Challenger. “There were two CapComs, the weather guy and the prime guy, and so it had been planned for some time that I’d be in the prime seat for [51L] and be the guy talking to them. . . . As the ascent Cap­Com you work so much with the crew that you have a lot of [connection]. In the training periods and stuff, not only do you sit over in the control center while they’re doing ascents and talk to them, but you also go and work with them on other things.”

Covey remembered getting together with the crew while the astronauts were in quarantine at jsc, before they flew down to Florida, to go over

To Touch the Face of God

31. Crew members of mission STS-51L stand in the White Room at Launchpad 39B. Left to right. Christa McAuliffe, Gregory Jarvis, Judy Resnik, Dick Scobee, Ronald McNair, Michael Smith, and Ellison Onizuka. Courtesy nasa.

the mission one more time and work through any questions. “We got to go over and spend an hour or two in the crew quarters with them. I spent most of my time with Mike Smith and Ellison Onizuka, who was my long­time friend from test pilot school. They were excited, and they were raun­chy, as you would expect, and we had a lot of fun and a lot of good laughs. It was neat to go do that. So that was the last time that I got to physical­ly go and sit with the crew and talk about the mission and the ascent and what to expect there.”

On launch day the flight control team reported much earlier than the crew, monitoring the weather and getting ready for communication checks with the astronauts once they were strapped in. Covey said that he was excited to be working with Flight Director Jay Greene, whom he had worked with be­fore, and that everything had seemed normal from his perspective leading into the launch. “From the control center standpoint,” Covey said, “I don’t remember anything that was unusual or extraordinary that we were working or talking about. It wasn’t something where we knew that someone was mak­ing a decision and how they were making that decision. We just flat didn’t have that insight. Didn’t know what was going on. Did not. It was pretty much just everything’s like a sim as we’re sitting there getting ready to go.”

Covey recalled that televisions had only recently been installed in the Mis­sion Control Center and that the controllers weren’t entirely sure what they were supposed to make of them yet. “The idea [had been] you shouldn’t be looking at pictures; You should be looking at your data,” he said. “So that’s how we trained. Since the last time I’d been in the control center, they’d started putting [televisions in]. . . . I’d sat as the weather guy, and once the launch happens, I kind of look at the data, but I look over there at the Tv.”

Astronaut Fred Gregory was the weather CapCom for the 51L launch and recalled that nothing had seemed unusual leading up to the launch.

Up to liftoff, everything was normal. We had normal communication with the crew. We knew it was a little chilly, a little cold down there, but the ice team had gone out and surveyed and had not discovered anything that would have been a hazard to the vehicle. Liftoff was normal. . . . Behind the flight director was a monitor, and so I was watching the displays, but also every now and then look over and look at Jay Greene and then glance at the monitor. And I saw what ap­peared to be the solid rocket booster motor’s explosive devices—what I thought— blew the solid rocket boosters away from the tank, and I was really surprised, because I’d never seen it with such resolution before, clarity before. Then I sud­denly realized that what I was intellectualizing was something that would oc­cur about a minute later, and I realized that a terrible thing had just happened.

Covey said Gregory’s reaction was the first indication he had that some­thing was wrong. “Fred is watching the video and sees the explosion, and he goes, ‘Wha—? What was that?’ Of course, I’m looking at my data, and the data freezes up pretty much. It just stopped. It was missing. So I look over and could not make heads or tails of what I was seeing, because I didn’t see it from a shuttle to a fireball. All I saw was a fireball. I had no idea what I was looking at. And Fred said, ‘It blew up,’ something like that.”

Covey recalled that the cameraman inside the control room continued to record what was happening there. “Amazingly, he’s still sitting there just crank­ing along in the control center while this was happening. Didn’t miss a beat,” he said, “because I’ve seen too many film footages of me looking in disbelief at this television monitor trying to figure out what the hell it was I was seeing.”

Off loop, Covey and Flight Director Jay Greene were talking, trying to gather information about what just happened. “There was a dialogue that started ensuing between Jay and myself,” Covey recalled, and Jay, he’s trying to get confirmation on anything from anybody, if they have any data, and what they think has happened, what the status of the or – biter is. All we could get is the solid rocket boosters are separated. Don’t know what else. I’m asking questions, because I want to tell the crew what to do. That’s what the ascent CapComs are trained to do, is tell them what to do. If we know something that they don’t, or we can figure it out faster, tell them so they can go and do whatever they need to do to recover or save themselves. There was not one piece of information that came forward; I was asking. I didn’t do it over the loop, so I did this between Jay and some of the other peo­ple that could hear, “Are we in a contingency abort? If so, what type of con­tingency abort? Can we confirm they’re off the SRBs?” Trying to see if there was anything I should say to the crew.

In all the confusion, he said, no one said anything about him attempt­ing to contact the crew members, since no one knew what to tell them. “We didn’t have any comm. We knew that. That was pretty clear to me; so the only transmissions that I could have made would have been over a UHF [ultrahigh frequency], but if I didn’t have anything to say to them, why call them? So we went through that for several minutes, and so if you go and look at it, there was never a transmission that I made after ‘ Challenger, you’re go [at] throttle up.’ That was the last one, and there wasn’t another one.”

After a few minutes of trying to figure out if there was anything to tell the crew, reality started to hit. Covey said,

I remember Jay finally saying, “Okay, lock the doors. Everybody, no commu­nications out. Lock the doors and go into our contingency modes of collecting data. ” I think when he did that, I finally realized; I went from being in this mode of, “What can we do? How do we figure out what we can do? What can we tell the crew? We’ve got to save them. We’ve got to help them save themselves. We’ve got to do something, ” to the realization that my friends had just died. … Of course, Fred and I were there together, which helped, because so many of the Challenger crew were our classmates, and so we were sharing that together. A special time that I’ll always remember being with Fred was there in the con­trol center for that.

It was a confusing time for those in the Launch Control Center. The data being received was not real-time data, Gregory said; there was a slight delay. “I had seen the accident occur on the monitor. I was watching data come in, but I saw the data then freeze, but I still heard the commentary about a normal flight coming from the public affairs person, who then, seconds later, stopped talking. So there was just kind of stunned silence in Mission Control.”

“At this point,” Gregory said, “no one had realized that we had lost the orbiter. Many, I’m sure, thought that this thing was still flying and that we had just lost radio signals with it. I think all of these things were kind of running through our minds in the first five to ten seconds, and then every­body realized what was going on.”

What Covey and Gregory, relatively insular in their flight control du­ties, did not realize was that concerns over the launch had begun the day before. The launch had already been delayed six times, and because of the significance of the first Teacher in Space flight and other factors, many were particularly eager to see the mission take off. On the afternoon of 27 Janu­ary (the nineteenth anniversary of the loss of astronauts Gus Grissom, Ed White, and Roger Chaffee in the Apollo 1 pad fire), discussions began as to whether the launch should be delayed again. The launch complex at Ken­nedy Space Center was experiencing a cold spell atypical for the Florida coast, with temperatures on launch day expected to drop down into the low twenties Fahrenheit in the morning and still be near freezing at launch time.

During the night, discussions were held about two major implications of the cold temperatures. The first was heavy ice buildup on the launch – pad and vehicle. The cold wind had combined with the supercooling of the cryogenic liquid oxygen and liquid hydrogen in the external tank to lead to the formation of ice. Concerns were raised that the ice could come off during flight and damage the vehicle, particularly the thermal protec­tion tiles on the orbiter. A team was assigned the task of assessing ice at the launch complex.

The second potential implication was more complicated. No shuttle had ever launched in temperatures below fifty degrees Fahrenheit before, and there were concerns about how the subfreezing temperatures would affect the vehicle, and in particular, the O-rings in the solid rocket boosters. The boosters each consisted of four solid-fuel segments in addition to the nose cone with the parachute recovery system and the motor nozzle. The seg­ments were assembled with rubberlike O-rings sealing the joints between

To Touch the Face of God

32. On the day of Space Shuttle Challengers 28 January 1986 launch, icicles draped the launch complex at the Kennedy Space Center. Courtesy nasa.

the segments. Each joint contained both a primary and a secondary O- ring for additional safety. Engineers were concerned that the cold temper­atures would cause the O-rings to harden such that they would not fully seal the joints, allowing the hot gasses in the motor to erode them. Burn – through erosion had occurred on previous shuttle flights, and while it had never caused significant problems, engineers believed that there was poten­tial for serious consequences.

During a teleconference held the afternoon before launch, srb con­tractor Morton Thiokol expressed concerns to officials at Marshall Space Flight Center and Kennedy Space Center about the situation. During a second teleconference later that evening, Marshall Space Flight Center of­ficials challenged a Thiokol recommendation that nasa not launch a shut­tle at temperatures below fifty-three degrees Fahrenheit. After a half-hour off-1 ine discussion, Thiokol reversed its recommendation and supported launch the next day. The three-hour teleconference ended after 11:00 p. m. (in Florida’s eastern time zone).

During discussions the next morning, after the crew was already aboard the vehicle, orbiter contractor Rockwell International expressed concern that the ice on the orbiter could come off during engine ignition and ric­ochet and damage the vehicle. The objection was speculation, since no launch had taken place in those conditions, and the NASA Mission Man­agement Team voted to proceed with the launch. The accident investiga­tion board later reported that the Mission Management Team members were informed of the concerns in such a way that they did not fully under­stand the recommendation.

Launch took place at 11:38 a. m. on 28 January. The three main engines ignited seconds earlier, at 11:37:53, and the solid rocket motors ignited at 11:38:00. In video of the launch, smoke can be seen coming from one of the aft joints of the starboard solid rocket booster at ignition. The primary O-ring failed to seal properly, and hot gasses burned through both the pri­mary and secondary O-rings shortly after ignition. However, residue from burned propellant temporarily sealed the joint. Three seconds later, there was no longer smoke visible near the joint.

Launch continued normally for the next half minute, but at thirty-seven seconds after solid rocket motor ignition, the orbiter passed through an area of high wind shear, the strongest series of wind shear events recorded thus far in the shuttle program. The worst of the wind shear was encountered at fifty-eight seconds into the launch, right as the vehicle was nearing “Max Q,” the period of the highest launch pressures, when the combination of velocity and air resistance is at its maximum. Within a second, video cap­tured a plume of flame coming from the starboard solid rocket booster in the joint where the smoke had been seen. It is believed that the wind shear broke the temporary seal, allowing the flame to escape. The plume rapid­ly became more intense, and internal pressure in the motor began drop­ping. The flame occurred in a location such that it quickly reached the ex­ternal fuel tank.

The gas escaping from the solid rocket booster was at a temperature around six thousand degrees Fahrenheit, and it began burning through the exterior of the external tank and the strut attaching the solid rocket boost­er to the tank. At sixty-four seconds into the launch, the flame grew stron­ger, indicating that it had caused a leak in the external tank and was now burning liquid hydrogen escaping from the aft tank of the external tank. Approximately two seconds later, telemetry indicated decreasing pressure from the tank.

At this time, in the vehicle and in Mission Control, the launch still ap­peared to be proceeding normally. Having made it through Max Q, the ve­hicle throttled its engines back up. At sixty-eight seconds, Covey informed the crew it was “Go at throttle up.” Commander Dick Scobee responded, “Roger, go at throttle up,” the last communication from the vehicle.

Two seconds later, the flame had burned through the attachment strut connecting the starboard srb and the external tank. The upper end of the booster was swinging on its strut and impacted with the external tank, rup­turing the liquid oxygen tank at the top of the external tank. An orange fireball appeared as the oxygen began leaking. At seventy-three seconds into the launch, the crew cabin recorder records Pilot Michael Smith on the in­tercom saying, “Uh-oh,” the last voice recording from Challenger.

While the fireball caused many to believe that the Space Shuttle had ex­ploded, such was not the case. The rupture caused the external tank to lose structural integrity, and at the high velocity and pressure it was experienc­ing, it quickly began disintegrating. The two solid rocket boosters, still fir­ing, disconnected from the shuttle stack and flew freely for another thirty – seven seconds. The orbiter, also now disconnected and knocked out of proper orientation by the disintegration of the external tank, began to be torn apart by the aerodynamic pressures. The orbiter rapidly broke apart over the ocean, with the crew cabin, one of the most solid parts of the ve­hicle, remaining largely intact until it made contact with the water.

All of that would eventually be revealed during the course of the acci­dent investigation. At Mission Control, by the time the doors were opened again, much was still unknown, according to Covey.

[We] had no idea what had happened, other than this big explosion. We didn’t know if it was an srb that exploded. I mean, that was what we thought. We always thought SRBs would explode like that, not a big fireball from the exter­nal tank propellants coming together. So then that set off a period then of just trying to deal with that and the fact that we had a whole bunch of spouses and families that had lost loved ones and trying to figure out how to deal with that.

The families were in Florida, and I remember, of course, the first thing I wanted to do was go spend a little time with my family, and we did that. But then we knew the families were coming back from Florida and out to Ellington [Field, Houston], so a lot of us went out there to just be there when they came back in. I remember it was raining. Generally they were keeping them isolat­ed, but a big crowd of us waiting for them, they loaded them up to come home. Then over the next several days most of the time we spent was trying to help the Onizukas in some way; being around. Helping them with their family as the families flew in and stuff like that.

After being in Mission Control for approximately twelve hours—half of that prior to launch and the rest in lockdown afterward, analyzing data, Gregory finally headed home.

The families had all been down at the Kennedy Space Center for the liftoff and they were coming back home. Dick Scobee, who was the commander, lived within a door or two of me. And when I got home, I actually preceded the families get­ting home; I remember that. They had the television remote facilities already set up outside of the Scobees’ house, and it was disturbing to me, and so I went over and, in fact, invited some of those [reporters] over to my house, and I just talked about absolutely nothing to get them away from the house, so that when June Scobee and the kids got back to the house, they wouldn’t have to go through this gauntlet.

The next few days, Gregory said, were spent protecting the crew’s fami­lies from prying eyes. “There was such a mess over there that Barbara and I took [Scobee’s] parents and just moved them into our house, and they must have stayed there for about four or five days. Then June Scobee, in fact, came over and stayed, and during that time is when she developed this concept for the Challenger Center. She always gives me credit for be­ing the one who encouraged her to pursue it, but that’s not true. She was going to do it, and it was the right thing to do.”

Gregory recalled spending time with the Scobees and the Onizukas and the Smiths, particularly Mike Smith’s children. “It was a tough time,” he said.

It was a horrible time, because I had spent a lot of time with Christa McAuliffe and [her backup] Barbara Morgan, and the reason was because I had teachers in my family. On my father’s side, about four or five generations; on my moth­er’s side, a couple of generations. My mother was elementary school, and my dad was more in the high school. But Christa and I and Barbara talked about how important it was, what she was doing, and then what she was going to do on orbit and how it would be translated down to the kids, but then what she was going to do once she returned. So it was traumatic for me, because not only had I lost these longtime friends, with Judy Resnik and Onizuka and Ron Mc­Nair and Scobee, and then Mike Smith, who was a class behind us, but I had lost this link to education when we lost Christa.

Astronaut Sally Ride was on a commercial airliner, flying back to Hous­ton, when the launch tragedy occurred. “It was the first launch that I hadn’t seen, either from inside the shuttle or from the Cape or live on television,” Ride recalled.

The pilot of the airline, who did not know that I was on the flight, made an announcement to the passengers, saying that there had been an accident on the Challenger. At the time, nobody knew whether the crew was okay; nobody knew what had happened. Thinking back on it, it’s unbelievable that the pilot made the announcement he made. It shows how profoundly the accident struck people. As soon as I heard, I pulled out my NASA badge and went up into the cockpit. They let me put on an extra pair of headsets to monitor the radio traf­fic to find out what had happened. We were only about a half hour outside of Houston; when we landed, I headed straight back to the Astronaut Office at jsc.

Payload Specialist Charlie Walker was returning home from a trip to San Diego, California, when the accident occurred.

I can remember having my bags packed and having the television on and search­ing for the station that was carrying the launch. As I remember it, all the sta­tions had the launch on; it was the Teacher in Space mission. So I watched the launch, and to this day, and even back then I was still aggravated with news services that would cover a launch up until about thirty seconds, forty-five sec-

onds, maybe one minute in flight, after Max Q, and then most of them would just cut the coverage. “Well, the launch has been successful. ” [I would think,] “You don’t know what you’re talking about. You’re only thirty seconds into this thing, and the roughest part is yet to happen. ”

And whatever network I was watching ended their coverage. “Well, looks like we’ve had a successful launch of the first teacher in space. ” And they go off to the programming, and it wasn’t but what, ten seconds later, and I’m about to pick my bags up and just about to turn off the television and go out my room door when I hear, “We interrupt this program again to bring you this announcement. It looks like something has happened. ” I can remember seeing the long-range tracker cameras following debris falling into the ocean, and I can remember going to my knees at that point and saying some prayers for the crew. Because I can remember the news reporter saying, “Well, we don’t know what has happened at this point. ” I thought, “Well, you don’t know what has happened in detail, but anybody that knows anything about it can tell that it was not at all good. ”

Mike Mullane was undergoing payload training with the rest of the 62A crew at Los Alamos Labs in New Mexico. “We were in a facility that didn’t have easy access to a Tv,” said Mullane.

We knew they were launching, and we wanted to watch it, and somebody finally got a television or we finally got to a room and they were able to finagle a way to get the television to work, and we watched the launch, and they dropped it away within probably thirty seconds of the launch, and we then started to turn back to our training. Somebody said, “Well, let’s see if they’re covering it further on one of the other channels, ” and started flipping channels, and then flipped it to a channel and there was the explosion, and we knew right then that the crew was lost and that something terrible had happened.

Mullane theorized that someone must have inadvertently activated the vehicle destruction system or a malfunction caused the flight termination system to go off. “I was certain of it,” Mullane said.

I mean, the rocket was flying perfectly, and then it just blew up. It just looked like it had been blown up from this dynamite. Shows how poor you can be as a witness to something like this, because that had absolutely nothing to do with it.

But it was terrible. Judy was killed on it. She was a close friend. There were four people from our group that were killed. It was a terrible time. Really as bad as it gets. It was like a scab or a wound that just never had an opportunity to heal because you had that trauma.

Astronaut Mary Cleave recalled two very different sets of reactions to the tragedy from the people around her in the Astronaut Office. “For the guys in the corps, when you’re in the test pilot business, you’re sort of a tough guy,” she said.

It’s a part of the job. It’s a lousy part of the job, but it’s part of the job. But I mean, the secretaries and everybody else were really upset, so we spent some time with them. Before my first flight, I had signed up. I basically told my family, “Hey, I might not be coming back." When we flew, it was the heaviest payload to orbit. We were already having nozzle problems. I think a lot of us under­stood that the system was really getting pushed, but that’s what we’d signed up to do. I think probably a lot of people in the corps weren’t as surprised as a lot of other people were. I did crew family escort afterwards. I was assigned to help when the families came down, as an escort at jsc when the president came in to do the memorial service. Jim Buchli was in charge of the group; they put a marine in charge of the honor guard. So I got to learn to be an escort from a marine, which was interesting. I learned how to open up doors. This was sort of like it doesn’t matter if you’re a girl or boy, there’s a certain way people need to be treated when they’re escorted. So I did that. That was interesting. And it was nice to think that you could help at that point.

Charlie Bolden had just returned to Earth ten days earlier from his first spaceflight, 61c. His crew was wrapping up postmission debriefing, he recalled, and it gathered with others in the Astronaut Office to watch Challenger launch. “That was the end of my first flight, and we were in heaven. We were celebrat­ing as much as anybody could celebrate,” he said. “We sat in the Astronaut Office, in the conference room with everybody else, to watch Challenger. No­body was comfortable because of all the ice on the launchpad and everything. I don’t think there were many of us who felt we should be flying that day, but what the heck. Everybody said, ‘Let’s go fly.’ And so we went and flew.”

Bolden thought the explosion was a premature separation of the solid rocket boosters; he expected to see the vehicle fly out of the smoke and per­form a return-to-launch-site abort. “We were looking for something good to come out of this, and nothing came out except these two solid rocket boosters going their own way.”

It took awhile, but it finally sunk in: the vehicle and the crew were lost. “We were just all stunned, just didn’t know what to do,” Bolden said. “By the end of the day we knew what had happened; we knew what had caused the accident. We didn’t know the details, but the launch photog­raphy showed us the puff of smoke coming out of the joint on the right – hand solid rocket booster. And the fact that they had argued about this the night before meant that there were people from [Morton] Thiokol who could say, ‘Let me tell you what happened. This is what we predict­ed would happen.’”

Bolden was the family escort for the family of 51L mission specialist Ron­ald McNair. Family escorts are chosen by crew members to be with the fam­ilies during launch activities and to be a support to families if something happened to the crew, as was the case with 51L. Much of Bolden’s time in the year after the incident was spent helping the McNair family, which in­cluded Ron’s wife and two children. “I sort of became a surrogate, if you will, for [McNair’s children] Joy and Reggie, and just trying to make sure that Cheryl [McNair] had whatever she needed and got places when she was supposed to be there. Because for them it was an interminable amount of time, I mean years, that they went through the postflight grieving pro­cess and memorial services and that kind of stuff.”

Bolden’s 61c crewmate Pinky Nelson was on his way to Minneapolis, Minnesota, for the premier of the imax movie The Dream Is Alive, which included footage from Nelson’s earlier mission, 41c. Nelson recalled hav­ing worked closely with the 51L crew, which Nelson said would be using the same “rinky-dink little camera” as his crew to observe Halley’s Comet.

I’d spent a bunch of time trying to teach Ellison [Onizuka] how to find Halley’s Comet in the sky. [Dick] Scobee and I were really close friends because of 41C, so “Scobe” and I had talked a lot about his kind of a “zoo crew, ” about his crew and all their trials and tribulations. He really wanted to get this mission flown and over with. So I talked to them the night before, actually, from down at the Cape and wished them good luck and all that, and then the accident happened while I was on the airplane to Minneapolis.

Nelson flew back to Houston from Minneapolis that afternoon, arriving around the same time that the families were arriving from Florida. Nelson and his wife, Susie, and astronaut Ox van Hoften and his wife convened at the Scobees’ home.

“The national press was just god-awful,” Nelson said.

I’ve never forgiven some of those folks. . . . I mean, it’s their job, but still— for their just callous, nasty behavior. We just spent a lot of time just kind of over at Scobee’s, trying to just be there and help out. I still can’t drink fla­vored coffee. That’s the only kind of coffee June had, vanilla bean brew or something. So whenever I smell that stuff, that’s always my memory of that, is having bad coffee at Scobes house, trying to just get their family through the time, just making time pass. We had to unplug the phones. The press was parked out in front of the house. It was a pretty bad time for all that. We went over and tried to do what we could with some of the other families. My kids had been good friends with Onizuka’s kids; they’re the same age. Lorna [Onizuka] was having just a really hard time. Everyone was trying to help out where we could.

Memorial services were beginning to be held for the lost crew members even as the agency was continuing with its search for the cockpit and the bodies of the lost crew. “It was terrible, going to the memorial services,” admitted Mullane.

It was one of those things that didn’t seem to end, because then they were looking for the cockpit out there. I personally thought, “Why are we doing this? Leave the cockpit down there. What are you going to learn from it?" Because by then they knew the SRB was the problem. . . . I remember thinking, “Why are we even looking for that cockpit? Just bury them at sea. Leave them there."I’m glad they did, though, because later I heard it was really shallow where that cockpit was. It was like, I don’t know, like eighty feet or something, which is too shal­low, because somebody eventually would have found it and pulled it up on a net or been diving on it or something. So it’s good that they did look for it. So you had these several weeks there, and then they bring that cockpit up, and then you have to repeat all the memorial services again, because now you have remains to bury. And then plus on top of that, you had the revelation that it wasn’t an accident; it was a colossal screwup. And you had that to deal with. So it was a miserable time, about as bad as I’ve ever lived in my life, were those months surrounding, months and years, really, surrounding the Challenger tragedy.

Astronaut Bryan O’Connor was at Kennedy Space Center during the debris recovery efforts and postrecovery analysis. O’Connor recalled be­ing on the pier when representatives from the Range Safety Office at Cape Canaveral were trying to determine whether what happened was an inad­vertent range safety destruct—if somehow there had been a malfunction of the destruct package intended to destroy the vehicle should a problem cause it to pose a safety risk to those on the ground. “I remember there was a Coast Guard cutter that came in and had some pieces and parts of the external tank,” O’Connor said. “On the second or third day, I think, one of these ships actually had a piece of the range safety destruct system from the external tank, intact for about halfway and then ripped up the other half of it. When he looked at that, he could tell that it hadn’t been a destruct.”

O’Connor had accident investigation training and was then assigned to work with Kennedy Space Center on setting up a place to reconstruct the ve­hicle as debris was recovered. “I remember we put tape down on the floor. We got a big room in the Logistics Center. They moved stuff out of the way. As time went on, the need increased for space, and we actually ended up putting some things outside the Logistics Center, like the main engines and some of the other things. But the orbiter pretty much was reassembled piece by piece over a period of time as the parts and pieces were salvaged out of the water, most of them floating debris, but some, I think, was picked up from subsurface.”

Recovery efforts started with just a few ships, O’Connor said, but grew into a large fleet. According to the official Rogers Commission Report on the accident, sixteen watercraft assisted in the recovery, including boats, subma­rines, and underwater robotic vehicles from NASA, the navy, and the air force. “It was one of the biggest salvage efforts ever, is what I heard at the time,” O’Connor said. “Over a period of time, we were able to rebuild quite a bit of the orbiter, laying it out on the floor and, in some cases, actually putting it in a vertical structure. Like the forward fuselage, for example, we tried to make a three-dimensional model from the pieces that we recovered there.”

While the goal had originally been to determine the cause of the acci­dent, the investigation eventually shifted to its effects, with analysis of the

To Touch the Face of God

33- This photograph, taken a few seconds after the loss of Challenger, shows the Space Shuttle’s main engines and solid rocket booster exhaust plumes entwined around a ball of gas from the external tank. Courtesy nasa.

debris revealing how the various parts of the vehicle had been affected by the pressures during its disintegration.

Astronaut Joe Kerwin, a medical doctor before his selection to the corps and a member of the first crew of the Skylab space station, was the direc­tor of Space Life Sciences at Johnson Space Center at the time of the acci­dent. “Like everybody else at jsc I remember exactly where I was when it happened,” Kerwin recalled.

I didn’t see it live. In fact, I was having a staff meeting in my office at jsc and we had a monitor in the background because the launch was taking place. And I just remember all of us sort of looking up and seeing this ex­plosion taking place on the monitor. And there was the moment of silence as each of us tried to absorb what it looked like was or might be going on, and then sort of saying, “Okay, guys, I think we better get to work. We’re going to need to coordinate with the Astronaut Office. We’re going to have to have flight surgeons. Sam, you contact the doctors on duty down at the

Cape and make sure that they have the families covered,” and we just sort of set off like that.

His medical team’s first actions were simply to take care of the families of the crew members, Kerwin said. “Then as the days went by and the search for the parts of the orbiter was underway I went down to Florida and coor­dinated a plan for receiving bodies and doing autopsies and things of that nature. It included getting the Armed Forces Institute of Pathology to com­mit to send a couple of experts down if and when we found remains to see whether they could determine the cause of death.”

But the crew compartment wasn’t found immediately and Kerwin went back home to Houston, where days turned into weeks.

I was beginning to almost hope that we wouldn’t have to go through that ex­cruciating investigation when I had a call from Bob Crippen that said, no, we’ve found the crew compartment and even at this late date there are going to be some remains so how about let’s get down here. I went down immediately.

By that time the public and press response to the accident and to NASA had turned bad, and NASA, which had always been considered one of the best or­ganizations in government, was now one of the worst organizations in govern­ment and there was a lot of bad press and there were a lot of paparazzi there in Florida who just wanted to get in on the action and get gruesome pictures or details or whatever they could. So we had to face that.

In addition to dealing with the press, Kerwin said recovery efforts also had to deal with local politics.

The local coroner was making noises like this accident had occurred in his ju­risdiction and therefore he wanted to take charge of any remains and perform the autopsies himself, which would have been a complicating factor, to say the least. I didn’t have to deal with that, I just knew about it and that I might have had to deal with [the] coroner if the offensive line didn’t block him. But the higher officials in NASA and in particular in the State of Florida got him called off saying, “No, this accident was in a federal spacecraft and it occurred offshore and you just back off”

By the time Kerwin arrived at the scene, the recovery team had come up with several different possible ways to get the remains to where they would be autopsied.

In view of the lateness of the time and in view of the press coverage and all that stuff, we decided that we needed a much more secure location for this activity to take place and we were given space in one of the hangars at Cape Canav­eral, one of the hangars in which the Mercury crews trained way back in the early sixties. We quickly prepared that space for the conduct of autopsies. When the remains were brought in they were brought into one of the piers by motor­ized boat. It was done after dark, and there was always one or two or three as­tronauts with the remains and we put up a little screen because right across the sound from this pier the press had set up floodlights and cameras and bleachers.

An unmarked vehicle was used to transport the remains to the hangar, while a nasa ambulance was used as a decoy to keep the location of the autopsies a secret. “We knew by that time, and really knew from the begin­ning, that crew actions or lack of actions didn’t have anything to do with the cause of the Challenger accident,” Kerwin said.

As an accident investigator you ask yourself that first—could this have been pilot error involved in any way, and the immediate answer was no. But we still needed to do our best to determine the cause of death, partly because of public interest but largely because of family interest in knowing when and under what circumstances their loved ones died. So that was the focus of the investigation.

The sort of fragmentary remains that were brought in, having been in the water for about six weeks, gave us no clue as to whether the cause of death was ocean impact or whether it could’ve taken place earlier. So the only thing left for the autopsies was to determine, was to prove, that each of the crew mem­bers had had remains recovered that could identify that, yes, that crew mem­ber died in that accident. That was not easy but we were able to do it. So at the end of the whole thing I was able to send a confidential letter to the next ofkin of each of the crew members stating that and stating what body parts had been recovered, a very, very short letter.

In addition to identifying the remains, Kerwin and his team worked to identify the exact cause of death of the crew members.

We’d all seen the breakup on television, we knew it was catastrophic, and my first impression as a doctor was that it probably killed all the crew just right at the time of the explosion. But as soon as they began to analyze the camera foot­age, plus what very little telemetry they had, it became apparent that the g-forces were not that high, not as high as you’d think. I guess the crew compartment was first flung upward and away from the exploding external tank and then rapidly decelerated by atmospheric pressure until it reached free fall. The explo­sion took place somewhere about forty thousand feet, I think about forty-six or forty-seven thousand, and the forces at breakup were estimated to be between fifteen and twenty gs, which is survivable, particularly if the crew is strapped in properly and so forth. The crew compartment then was in free fall, but up­ward. It peaked at about sixty-five thousand feet, and about two and half min­utes after breakup hit the ocean at a very high rate of speed. If the crew hadn’t died before, they certainly died then.

Kerwin said he and his team then worked to refine their understand­ing of exactly what had happened when. “Our investigation attempted to determine whether or not the crew compartment’s pressure integrity was breeched by the separation, and if so, if the crew compartment had lost pressure, we could then postulate that the crew had become profoundly un­conscious because the time above forty thousand feet was long enough and the portable emergency airpack was just that, it was an airpack, not an ox­ygen pack, so they had no oxygen available. They would have become un­conscious in a matter of thirty seconds or less depending on the rate of the depressurization and they would have remained unconscious at impact.”

But the damage to the compartment was too great to allow Kerwin to determine with certainty whether the cabin had lost pressure. He said that, given what it would mean for the crew’s awareness of its fate, his team was almost hopeful of finding evidence that the crew compartment had been breached, but they were unable to make a conclusive determination.

The damage done to the crew compartment at water impact was so great that despite a really lot of effort, most of it pretty expert, to determine whether there was a pressure broach in any of the walls, in any of the feed-throughs, any of the windows, any of the weak points where you might expect it, we couldn’t rule it out but we couldn’t demonstrate it either. We lifted all the equipment to see whether any of the stuff in the crew compartment looked as if it had been dam­aged by rapid decompression, looked at toothpaste tubes and things like that, tested some of them or similar items in vacuum chambers to determine wheth­er that sort of damage was pressure-caused and that was another blind hole. We simply could not determine. And then as to the remains, in a case of immedi­ate return of the remains we might have gotten lucky and been able to measure tissue oxygen concentration and tissue carbon dioxide concentration and gotten a clue from that but this was way too late for that kind of thing.

Finally, after all ofthat we had to just, I just had to sit down and write the letter to Admiral [Richard H.] Truly stating that we did not know the cause of death.

Sally Ride was named to the commission that was appointed to in­vestigate the causes of the accident. President Reagan appointed former secretary of state William Rogers to head a board to determine the fac­tors that had resulted in the loss of Challenger. (Ride has the distinction of being the only person to serve on the presidential commission inves­tigations of both the Challenger and the Columbia accidents.) Ride was in training at the time of the accident and was assigned to the presiden­tial commission within just a few days of the accident. The investigation lasted six months.

“The panel, by and large, functioned as a unit,” Ride recalled.

We held hearings; we jointly decided what we should look into, what witness­es should be called before the panel, and where the hearings should be held. We had a large staff so that we could do our own investigative work and conduct our own interviews. The commission worked extensively with the staff through­out the investigation. There was also a large apparatus put in place at NASA to help with the investigation: to analyze data, to look at telemetry, to look through the photographic record, and sift through several years of engineering records. There was a lot of work being done at NASA under our direction that was then brought forward to the panel. I participated in all ofthat. I also chaired a sub­committee on operations that looked into some of the other aspects of the shut­tle flights, like was the astronaut training adequate? But most of our time was spent on uncovering the root cause of the accident and the associated organiza­tional and cultural factors that contributed to the accident.

Ride said she had planned to leave nasa after her upcoming third flight and do research in an academic setting. But her role in the accident in­vestigation caused her to change her plans. “I decided to stay at nasa for an extra year, simply because it was a bad time to leave,” explained Ride.

She described that additional year as one that was very difficult both for her personally and for the corps. “It was a difficult time for me and a dif­ficult time for all the other astronauts, for all the reasons that you might expect,” said Ride. “I didn’t really think about it at the time. I was just go­ing from day to day and just grinding through all the data that we had to grind through. It was very, very hard on all of us. You could see it in our faces in the months that followed the accident. Because I was on the com­mission, I was on Tv relatively frequently. They televised our hearings and our visits to the NASA centers. I looked tired and just kind of gray in the face throughout the months following the accident.”

Astronaut Steve Hawley, who was married to Ride at the time, recalled providing support to the Rogers Commission, particularly in putting to­gether the commission’s report, Leadership and America’s Future in Space.

The chairman felt that it was appropriate to look not only at the specifics of that accident but other things that his group might want to say about safety in the program, and that included, among other things, the role of astronauts in the program, and that was one of the places where I think I contributed, was how astronauts ought to be involved in the program. I remember one of the recom­mendations of the committee talked about elevating the position of director of FCOD [Flight Crew Operations Directorate], because at the time of the Chal­lenger accident, he was not a direct report to the center director. That had been a change that had been made sometime earlier, I don’t remember exactly when. And that commission felt that the guy that was head of the organization with the astronauts should be a direct report to the center director. So several people went back to Houston and put George [Abbey] s desk up on blocks in an at­tempt to elevate his position. I think he left it there for some period of time, as far as I can remember.

Though not everyone in the Astronaut Office was involved in the Rog­ers Commission investigation and report, many astronauts were involved in looking into potential problems in the shuttle hardware and program. Recalled O’Connor,

I remember that a few days after the accident Dick Truly and the acting ad­ministrator [William Graham] were down at ksc. Again, because I had ac­cident investigation training, they had a discussion about what’s the next steps

here. The acting administrator, it turned out, was very reluctant to put a board together, a formal board. We had an Accident Investigation Team that was as­sembling; eventually became under the cognizance of J. R. Thompson. We had a bunch of subteams under him, and then he was reporting to Dick Truly. But we never called out a formal board, and I think there was more or less a politi­cal feel that this is such high visibility that we know for sure that the Congress or the president or both are going to want to have some sort of independent in­vestigation here, so let’s not make it look like we’re trying to investigate our own mishap here. And that’s why he decided not to do what our policy said, which is to create a board. We stopped short of that.

Once the presidential commission was in place, O’Connor said it was obvious to Dick Truly that he needed to provide a good interface between the commission and nasa. Truly assigned O’Connor to set up an “action center” in Washington DC for that purpose. The center kept track of all the requests from the commission and the status of the requests.

“They created a room for me, cleared out all the desks and so on,” O’Connor recalled. “We put [up] a bunch of status boards; very old-fashioned by to­day’s standard, when I think about it. It was more like World War Il’s tech­nology. We had chalkboards. We had a paper tracking system, an IBM type­writer in there, and so on. It all seemed so ancient by today’s standards, virtually nothing electronic. But it was a tracking system for all the requests that the board had. It was a place where people could come and see what the status of the investigation was.”

The center became data central for nasa’s role in the investigation. In addition to the data keeping done there, Thompson would have daily tele­conferences with all of his team members to find out what was going on.

Everybody would report in, what they had done, where they were, where they were on the fault tree analysis that we were doing to x out various potential cause factors. I remember the action center became more than just a place where we coordinated between NASA and the blue-ribbon panel. It was also a place where people could come from the [Capitol] Hill or the White House. We had quite a few visitors that came, and Dick Truly would bring them down to the action center to show them where we were in the investigation. So it had to kind of take on that role, too, of publicly accessible communication device.

I did that job for some weeks, and then we rotated people. [Truly] asked that George Abbey continue [to provide astronauts]. George had people available now that we’re not going to fly anytime soon, so he offered a bunch of high-quality people. . . . So shortly after I was relieved from that, I basically got out of the investigation role and into the “what are we going to do about it" role and was assigned by George to the Shuttle Program Office.

As board results came out, changes began to be made. For example, O’Connor said, all of the mission manifests from before the accident were officially scrapped.

I was relieved of my job on a crew right away. I think they called it 6im, which was a mission I was assigned to right after I got back from 6ib. I cant remem­ber who all the members were, but my office mate, Sally Ride, and I were both assigned to that same mission, 6im. Of course, when the accident happened, all that stuffbecame questionable and we stopped training altogether. I don’t think that mission ever resurrected. It may have with some other name or number, but the crew was totally redone later, and the two of us got other assignments.

In the meantime, O’Connor was appointed to serve as the assistant pro­gram manager for operations and safety. His job included coordinating how NASA was going to respond to a couple of the major recommendations that came out of the blue-ribbon panel. The panel had ten recommendations, covering a variety of subjects. One of them had to do with how to restruc­ture and organize the safety program at NASA. Another one O’Connor was involved with dealt with wheels, tires, brakes, and nose-wheel steering.

“It was all the landing systems,” he explained.

Now, that may sound strange, because that had nothing to do with this accident, but the Challenger blue-ribbon panel saw that, as they were looking at our his­tory on shuttle, they saw that one of the bigger problems we were addressing tech­nically with that vehicle was landing rollout. We had a series of cases where we had broken up the brakes on rollout by overheating them or overstressing them. We had some concerns about automatic landings. We had some concerns about steering on the runway in cases of a blown tire or something like that. So they chose to recommend that we do something about these things, put more empha­sis on it, make some changes and upgrades in that area.

In the wake of the accident, even relatively low-profile aspects of the shuttle program were reexamined and reevaluated. “nasa wanted to go back and look at everything, not just the solid rocket boosters, but every­thing, to determine, is there another Challenger awaiting us in some oth­er system,” recalled Mullane, who was assigned to review the range safety flight termination system, which would destroy the vehicle in the event it endangered lives. “I always felt it was a moral issue on this dynamite sys­tem. I always felt it was necessary to have that on there, because your wives and your family, your Lcc [Launch Control Center] people are sitting there two and a half miles away. If you die, that’s one thing. But if in the process of you dying that rocket lands on the LCC and kills a couple of hundred people, that’s not right. So they should have dynamite aboard it to blow it up in case it is threatening the civilian population.”

While Mullane himself supported having the flight termination system on the vehicles, he said that some of his superiors did not.

I could not go to the meetings and present their position. I couldn’t. I mean, to me it was immoral. It was immoral to sit there and say we fly without a dynamite system aboard. That’s immoral. And we threatened lcc, we threat­en our families, we threaten other people. We’re signing up for the risk to ride in the rocket.

That was another bad time of my life, because I took a position that was counter to my superiors’ position, and I felt that it was jeopardizing my future at nasa. I didn’t like that at all, didn’t like the idea that I was supposed to just parrot somebody’s opinion and mine didn’t count on that issue. . . . So I did not like that time of my life at all. I had an astronaut come to me once. . . who heard my name being kicked around as a person that was causing some prob­lems. And that’s the last thing you want in your career is to hear that your name is in front of people who make launch crew decisions, who make crew decisions, and basically, it looks like I’m a bad apple. But I just couldn’t do it. . . . I said, “You go. You do it. I cant. It’s immoral." By the way, the end of that is that the solid boosters retained their dynamite system aboard. It was taken off the gas tank, making it much safer, at least now two minutes up when the boosters are gone, we don’t have to dynamite aboard anymore, so it could fail and blow you up. But that’s the right decision right there. You protect the civilians, you pro­tect your family, you protect lcc with that system.

The results from the commission questioned the NASA culture, and ac­cording to Covey some within the agency found that hard to deal with, es­pecially while still recovering from the loss of the crew and vehicle. “Every­body was reacting basically to two things,” Covey said. “One was the fact that they had lost a Space Shuttle and lost a crew, and two, the Rogers Com­mission was extremely critical, and in many cases, rightfully so, about the way the decision-making processes had evolved and the culture had evolved.”

Covey noted,

Those two things together are hard for any institution to accept, because this was still a largely predominant workforce that had come through the Apollo era into the shuttle era and had been immensely successful in dealing with the issues that had come through both those programs to that point. So to be told that the culture was broken was hard to deal with, and that’s because culture doesn’t change overnight, and there was a lot of people that didn’t believe that that was an accurate depiction of the situation and environment that existed within the agency, particularly at the Johnson Space Center.

Personnel changes began to take place, including the departures of George Abbey, the long-time director of flight crew operations, and John Young, chief of the Astronaut Office. “We started seeing a lot of personnel chang­es in that time period in leadership positions,” Covey said.

I think they were a matter oftiming and other things, but George Abbey had been the director of flight crew operations for a long time, and somewhere in there George left that position. Don Puddy came in as the director of flight crew operations, and that sat poorly with a lot of people, because he wasn’t out of Flight Crew Opera­tions; he was a flight director. So that was something that a lot of people just had a hard time accepting. John Young left from being chief of the Astronaut Office, and Dan Brandenstein came into that position. I’m trying to remember what hap­pened in Mission Operations, but somewhere in there Gene Kranz left; I cant re­member when, and I’m not sure when in that spectrum of things that he did. So there were changes there. The center director changed immediately after the Chal­lenger accident, also, and changes were rampant at headquarters. So, basically, there was a restructuring of the leadership team from the administrator on down.

In time, the astronaut corps grieved and adjusted to leadership changes, and their focus returned to flying. “As jsc does,” Covey said, “once they got past the grief and got past the disappointment of failure and acceptance of their role relative to the decision-making process, they jumped in and said, ‘Okay, our job is flying. Let’s go figure out how we’re going to fly again.’”

Covey said the Challenger accident caused some in the astronaut corps to be wary of the NASA leadership structure and to distrust that the system would make the right kinds of decisions to protect them. “I wouldn’t say it was a bunker mentality, but it was close to that,” Covey said.

The idea that, as it played out, that there were decisions made and informa­tion that may not have been fully considered, and as you can see from all of it, a relatively limited involvement of any astronauts or flight crew people in the decision that led up to the launch; very little, if any. So that led to some changes that have evolved over time where there are more and more astronauts that have been involved in that decision-making process at the highest levels, either within the Space Shuttle program or in the related activities, where be­fore it was like, “Yeah, those guys will make the right decision, and we’ll go fly."

Along those same lines, Fred Gregory commented that the tragedy was a moment of realization for many about the dangers of spaceflight.

The first four missions we called test flights, and then on the fifth mission we de­clared ourselves operational. We were thinking offlying journalists, and we had Pete Aldridge, who was the secretary of the air force. [about to fly]. I mean, we were thinking of ourselves as almost like an airline at that point. It came back safely. Everything was okay, even though there may have been multiple failures or things that had degraded. When we looked back, we saw that, in fact, we had had this erosion of those primary and secondary O-rings, but since it was a successful mission and we came back, it was dismissed almost summarily. I think there was a realization that we were vulnerable, and that this was not an airliner, flying to space was risky, and that we were going to have to change the approach that we had taken in the past.

During the investigation, new light was shined on past safety issues and close calls. Astronaut Don Lind recalled being informed of a situation that occurred on his 51B mission, less than a year earlier.

What happened was that Bob Overmyer and I had shared an office for three and a half years, getting ready for this mission. Bob, when they first started the

Challenger investigation, was the senior astronaut on the investigation team. He came back from the Cape one day, walked in the office, slumped down in the chair, and said, “Don, shut the door." Now, in the Astronaut Office, if you shut the door, it’s a big deal, because we tried to keep an open office so people could wander in and share ideas. . . . So I shut the door, and Bob said, “The board today found out that on our flight nine months previously we almost had the same explosion. We had the same problems with the O-rings on one of our boosters." We talked about that for a few minutes, and he said, “The board thinks we came within fifteen seconds of an explosion. ”

After learning this, Lind traveled out to what was then Morton Thio – kol’s solid rocket booster facility in Utah to find out exactly what had hap­pened on his flight.

Alan McDonald, who was the head of the Booster Division, sat down with me. . . . He got out his original briefing notes for Congress, which I now have, and outlined exactly what had happened. There are three separate O-rings to seal the big long tubes with the gases flowing down through them at about five thousand degrees and 120 psi. The first two seals on our flight had been total­ly destroyed, and the third seal had 24 percent of its diameter burned away. McDonald said, “All of that destruction happened in six hundred millisec­onds, and what was left of that last O-ring if it had not sealed the crack and stopped that outflow of gases, if it had not done that in the next two hundred to three hundred milliseconds, it would have been gone all the way. You’d nev­er have stopped it, and you’d have exploded. So you didn’t come within fifteen seconds of dying, you came within three-tenths of one second of dying. " That was thought provoking.

In the wake of the Challenger disaster, all future missions were scrubbed and the flight manifest was replanned. Some missions were rescheduled with new numbers and only minor changes, while other planned missions and payloads were canceled entirely.

Recalled Charlie Walker of the electrophoresis research he had been con­ducting, “The opportunity just went away with the national policy changes; commercial was fourth priority, if it was a priority at all, for shuttle man­ifest. . . . Shuttle would not be flying with the regularity or the frequency that had been expected before.”

The plans to use the Centaur upper stage to launch the Ulysses and Gal­ileo spacecraft from the shuttle were also canceled as a result of the acci­dent. Astronaut Mike Lounge had been assigned to the Ulysses Centaur flight, which was to have flown in spring 1986 on Challenger. “So when we saw Challenger explode on January 28, before that lifted off, I remember thinking, ‘Well, Scobee, take care of that spaceship, because we need it in a couple of months.’ We would have been on the next flight of Challenger"

Lounge recalled that his crew was involved in planning for the mis­sion during the Challenger launch. They took a break from discussing the risky Centaur mission, including ways to eject the booster if necessary, so that they could watch the 51L launch on a monitor in the meeting room. Lounge said the disaster shone a new light on the discussions they’d been having. “We assumed we could solve all these problems. We were still ba­sically bulletproof. Until Challenger, we just thought we were bulletproof and the things would always work.”

Rick Hauck, who was assigned to command the Ulysses Centaur mission, was glad to see the mission canceled. “Would it have gotten to the point where I would have stood up and said, ‘This is too unsafe. I’m not going to do it’?” Hauck said. “I don’t know. But we were certainly approaching lev­els of risk that I had not seen before.”

For Bob Crippen, the Challenger accident would result in another loss, that of his final opportunity for a shuttle flight. The ramifications of the ac­cident meant that he ended his career as a flight-status astronaut the same way he began it—in the crew quarters for Vandenberg’s SLC-6 launch com­plex, waiting for a launch that would never come. In the 1960s that launch had been of the Air Force’s Manned Orbiting Laboratory. In 1986 it was shuttle mission 62A.

“If I have one flying regret in my life, it was that I never had an oppor­tunity to do that Vandenberg mission,” Crippen said.

We [had been] going to use filament-wound solid rockets, whereas the solid rock­ets that we fly on board the shuttle have steel cases. That was one of the things, I think, that made a lot of people nervous. . . . We needed the filament-wounds to get the performance, the thrust-to-weight ratio that we needed flying out of Vandenberg. So they used the filament-wound to take the weight out. After we had the joint problem on the solids with Challenger; most people just couldn’t get comfortable with the filament-wound case, so that was one of the aspects of why they ended up canceling it.

As the agency, along with the nation, was reevaluating what the future of human spaceflight would look like, what things should be continued, and what things should be canceled, the astronauts in the corps had to make the same sorts of decisions themselves. Many decided that the years-long gap in flights offered a good time to leave to pursue other interests, but many stayed on to continue flying.

“I think it was a very sobering time,” Jerry Ross said. “I mean, we always knew that that was a possibility, that we would have such a catastrophic ac­cident. I think there was a lot of frustration that we found out fairly soon afterwards that the finger was being pointed at the joint design and, in fact, that there had been quite a bit of evidence prior to the accident that that joint design was not totally satisfactory. Most of us had never heard that. We were very shocked, disappointed, mad.”

Several astronauts left the office relatively soon after the accident, Ross said, and a stream of others continued to leave for a while afterward. “There were more people that left for various different reasons; some of them for frustration, some of them knowing that the preparations for flight were go­ing to take longer, some of them responding to spouses’ dictates or requests that they leave the program now that they’d actually had an accident.”

Ross discussed his future at NASA with his family and admitted to hav­ing some uncertainty at first as to what he wanted to do.

I had mixed feelings at first about wanting to continue to take the risk of fly­ing in space, but at the same time, all of the NASA crew members on the shuttle were good friends of mine, and I felt that if I were to quit and everybody else were to quit, then they would have lost their lives for no good benefit or progress. If you reflect back on history, any great undertaking has had losses; you know, wagon trains going across the plains or ships coming across the ocean. I was just watching a TV show that said in the 1800s, one out of every six ships that went across. . . the ocean from Europe to here didn’t make it. So there’s risk involved in any type of new endeavor that’s going on. And I got into the program with my eyes wide open, both for the excitement and the adventure of it, but also I felt very strongly that it was important that we do those kinds of things for the future of mankind and for the good of America.

After getting through the shock and getting through the memorial services and all that, even though I was frustrated I wasn’t getting a whole lot to do to help with the recovery effort, I was very determined that I was not going to leave, after talking with the family and getting their agreement, and that I was going to do whatever I could to help us get back to flying as soon as we could and to do it safer.

Bold They Rise

After John Young and I made the first flight of the Space Shuttle aboard Co­lumbia all those years ago, people would sometimes ask me what the best part of the flight was. I would always use John’s classic answer: “The part between takeoff and landing.”

Now that it’s all said and done, I think that describes what the best part of the Space Shuttle program was: the part between our first launch in April 1981 and the last landing in July 2011.

There were some low points in between, particularly the loss of both of the orbiters I had the privilege to fly and their crews, but as a whole I think the shuttle has been one of the most marvelous vehicles that has ever gone into space—a fantastic vehicle unlike anything that’s ever been built.

The Space Shuttle has carried hundreds of people into space and deliv­ered hundreds of tons of payloads into space. The shuttle gave us the Gali­leo and Magellan probes, which opened our eyes to new worlds, and it let us not only launch the Hubble Space Telescope but also repair and upgrade it time and time again, and Hubble has revolutionized our understanding of not only our solar system but the entire universe. The shuttle carried a lot of classified military payloads early on that probably helped the United States win the Cold War.

The Space Shuttle let us build the International Space Station. The Space Station is an incredible accomplishment, a marvelous complex, but it was the Space Shuttle that taught us that we could build a complicated space vehicle and make it work very well. The Space Station would not have been possible without the Space Shuttle.

But in those early days, I think the shuttle did something else, a little less concrete but just as important. The late ’70s and early ’80s weren’t re­ally a great time for the United States. We’d basically lost the Vietnam War. We’d been through economic hard times, through the hostage crisis in Iran.

President Reagan was shot just before our flight on STS-i. And morale for a lot of people in the country was really low. People were feeling like things just weren’t going right for us.

And that first flight, it was obvious that it was a big deal. It was a big thing for NASA, but it was a big thing for the country. It wasn’t just our ac­complishment at NASA; it was an American accomplishment. It was a mo­rale booster for the United States. It was a rallying point for the American people. And the awareness may not be as high now as it was then, but I think that’s still true today. I think you saw that when the shuttle made its last flight; the pride people had in what it had accomplished and the fact that a million people watched it. When I talk to people, they think space exploration is something we need to be doing, for the future of the United States and humankind.

The retirement of the shuttle was kind of bittersweet for me. I’m proud of all it’s accomplished, and I’m sorry to see it end. But I believe in moving on. I’d like to see us get out of Earth orbit and go back to the moon, and to other destinations, and eventually to Mars.

John and I got to see a lot of the development of the Space Shuttle first­hand. As astronauts, we were involved from an operations standpoint, and as the first crew, John and I visited the sites where they were working on the shuttle, getting it ready to fly. We had an outstanding, dedicated team, people who really believed they were doing something important for the nation. When we finally got into the shuttle for that first flight, meeting those thousands of people gave me a lot of confidence that we had a good vehicle to fly on.

I never expected to be selected for that first flight. I thought they would pick someone more experienced to fly with John. I was excited that they picked me, and I was honored to be a part of that flight. All told, that flight was the beginning of something truly amazing, and I’m honored to be one of the thousands of people who made it happen.

Bob Crippen

Preface

When I (David) first became involved in the Outward Odyssey series, working on the Skylab volume, my coauthors and I were shown a list of proposed titles for the first eight books in the series. As authors working on our first book, coming up with a title seemed like one of the more ex­citing parts of the job. We were thus somewhat pleased to be disappointed with the working title the publisher had provided: “Exemplary Outpost.” It was an accurate title, but it lacked the poetry of the other titles on the list—titles like Into That Silent Sea and In the Shadow ofthe Moon. I’m not sure that we quite lived up to that standard with Homesteading Space, but we made our best effort.

Even though it meant giving up the privilege of titling this volume, Heather and I were quite happy to go along with the name the publisher had suggested for this book: Bold They Rise. It was, quite literally, poetic, taken from the poem by series editor Colin Burgess that appears as the epigraph.

When we first read the poem, very early on in the process of writing this volume, we pictured the title as being about the Space Shuttles themselves, reflecting the poem’s reference to “winged emissaries.” As the book took shape, however, we realized that was no longer true; the title had taken on a new meaning for us. Rather than being about the hardware, it was about the men and women who risked their lives to expand humankind’s frontiers.

And in that vein, this book owes an incredible debt of gratitude to the NASA Johnson Space Center (jsc) Oral History Project, without which it quite literally would not exist.

With Homesteading Space, it was relatively easy to create a book that filled a unique niche—with a few notable exceptions, such as a handful of official NASA publications and David Shayler’s Skylab, very little had been written about America’s first space station. Breaking new ground was not a particular challenge.

With this book, the challenge was a little greater. There are more books about the Space Shuttle program, so it was somewhat harder to create some­thing unique. Most of the previous works, however, fall into one of three categories—technical volumes, which span the entire program but include none of the human experience; astronaut memoirs, which relate the hu­man experience, but only from one person’s perspective; or specific histories, which are more exhaustive but focus on only a limited slice of the program.

Based on the overall goal of the Outward Odyssey series, a new niche we could address became clear—a book relating the human experience of the Space Shuttle program, not limited to one person’s story but including a variety of viewpoints and spanning the early years of the program. Origi­nally the goal was to create a “Homesteading Space of the shuttle program,” but it quickly became apparent that was a misdirected goal. Homesteading had only three manned missions to cover, and thus we could delve much deeper and more broadly in covering them. To attempt to write about the subject of this book in that manner would be to do either the subject or the reader a grave disservice; we needed to narrow our approach to create something that was both relevant and readable.

When we began reading from the jsc oral history interviews early in our research, the ideal approach for the book became apparent. Here was a wealth of first-person experience, describing in detail what it was like to be there—what it was like to involved in the design of a new spacecraft, what it was like to risk one’s life testing that vehicle, what it was like to do things that no one had done before in space, what it was like to float freely in the vacuum of space as a one-man satellite, what it was like to hold thousands of pounds of hardware in one’s hands, what it was like to watch friends die.

This book almost exclusively offers the astronauts’ perspective on the early years of the Space Shuttle program, and, while research for the vol­ume drew on several resources, the extensive quoted material draws heavi­ly from the jsc Oral History Project. It’s the astronauts’ story, told in their own words, about their own experiences.

Bold They Rise is not a technical volume. We would love for this volume to inspire you seek out another book that delves more deeply into the tech­nical aspects of the shuttle. There are parts of the story that we had to deal with in what seemed like a relatively superficial manner; even dedicating an entire chapter to the Challenger accident and the effects it had seems woeful­ly insufficient. Entire books could, and have, been written about the Chal­lenger accident. If this book leaves you wanting to know more about that incident or other aspects of the shuttle’s history, we encourage you to seek out those volumes. And of course, individual astronauts have told their sto­ries in memoirs with more personality than we were able to capture here. The subject of this book is such that it can’t be covered by any one volume exhaustively, but hopefully we have provided a unique, informative, and engaging overview here.

The chronological scope of the book was also set by the publisher to fit within the Outward Odyssey series. (Another volume, written by Rick Houston, picks up the Space Shuttle story where this one leaves off.) Ini­tially, the ending point of the book was a bit discomfiting; the Challenger accident seemed a rather low note on which to end a book. There were any number of successes both before and after Challenger. Why would one pick the lowest point of the early years as a place to end the story? But, in a very real way, it was the best possible way to turn this history into a story arc.

As astronaut Mike Mullane wrote in his memoir Riding Rockets,

The NASA team responsible for the design of the Space Shuttle was the same team that had put twelve Americans on the Moon and returned them safely to Earth across a quarter million miles of space. The Apollo program represented the greatest engineering achievement in the history of humanity. Nothing else, from the Pyramids to the Manhattan Project, comes remotely close. The men and women who were responsible for the glory of Apollo had to have been af­fected by their success. While no member of the Shuttle design team would have ever made the blasphemous claim, “We’re gods. We can do anything," the reality was this: The Space Shuttle itself was such a statement. Mere mortals might not be able to design and safely operate a reusable spacecraft boosted by the world’s largest, segmented, uncontrollable solid-fueled rockets, but gods certainly could.

That, then, is the story of this book—a Greek tragedy about hubris and its price. It’s a story of the confidence that bred some of the most amazing achievements in human history but also led to overconfidence.

But make no mistake, this book is also a love letter. Both authors of this volume were born after the end of the last Saturn-Apollo flight; the Space Shuttle is “our” spacecraft. The Challenger accident occurred when we were still children; it was our “where were you” equivalent of the Kennedy assas­sination. In our “day jobs” as NASA education writers, we wrote extensively about the shuttle, its crews, its missions, its accomplishment and ultimate­ly its retirement. We write this with a fondness for the shuttle, even when that means telling the story with warts-and-all honesty.

It’s been an honor and a pleasure to tell this story. We hope you enjoy reading it.

David Hitt

Heather R. Smith

Acknowledgments

As mentioned in the preface but bears repeating, this volume owes a great deal of gratitude to the Johnson Space Center Oral History Project, with­out which it would not exist.

In addition, we are grateful to the University of Nebraska Press, and in particular to senior editor Rob Taylor, for their dedication to chronicling the history of space exploration through their publication of the Outward Odyssey series and specifically through their help and support with this vol­ume. In addition, the authors wish to express their substantial thanks to Outward Odyssey series editor Colin Burgess, who has been a loyal shep­herd, a wise counsel, and a good friend during the process.

It was an incredible honor to have astronaut Bob Crippen agree to write the foreword for this volume. For David, the journey to writing this book begins in a very real way in front of a television set in 1981 watching Bob Crippen and John Young make history, and to conclude that journey with Crippen being a part of this project is a surreal bookend to the experience.

Astronaut (and Homesteading Space coauthor) Owen Garriott provided much assistance early in the project, making contacts and helping to get things moving, and that assistance is much appreciated. In addition, astro­naut Bo Bobko was also involved in the early stages of the book and pro­vided insight into its direction and helped open some doors. Astronauts Hank Hartsfield and Joe Kerwin and nasa legends Chris Kraft and George Mueller also provided us with material for the book.

Phillip Fox, Jon Meek, Jordan Walker, Rebecca Freeman, Lauren McPher­son, and Suzanne Haggerty read early portions of this book in progress and pro­vided feedback.

On a personal note, the authors wish to acknowledge Finn and Caden Smith, ages seven and five at the time the original manuscript was finished, for their sacrifices during deadline work on this book.

In addition, David would like to thank the following:

Heather, who for years has made my writing better and without whom I could not have written this book.

As per last time, my father, Bill Hitt, for engendering my interest in spaceflight that set me on the path to, among other things, writing this book. Jim Abbott, for giving me my first break and being a brilliant editor and a wonderful mentor and for shaping the man I am today. Holly Snow, for opening the door for my new involvement with nasa.

Owen Garriott and Joe Kerwin, for sponsoring me through Olympus and for sharing their stories, their insight, their knowledge, their expertise, and their friendship.

All of those who traveled with me on multiple road trips to Kennedy Space Center, which occasionally involved successfully watching shuttle launches.

Heather would like also to thank the following:

David, for offering me the opportunity to coauthor a book and for shep­herding me through the process.

Mrs. Hughes, for seeing potential in the writing skills of a young, tenth – grade Heather and inviting her to write for the school yearbook staff, spark­ing an interest in writing and communication that led me down this career path. Mr. Sandy Barnard, for believing that I could write and write well whatever I put my pen to.

The Times-Mail in Lawrence County, Indiana, the proud home to three astronauts, including Charlie Walker, who is quoted extensively in this book, for giving me my first professional writing job and an occasional space-related assignment that made a big difference in me ending up writ­ing at NASA and thus ending up writing this book. I was blessed to work in a community that adores its hometown astronauts and that still gets ex­cited about spaceflight.

Starbucks locations in Huntsville, Alabama, and Nashville, Tennessee, and the Flint River Coffee Company in Huntsville, Alabama, for hospi­tality and tasty coffee. Portions of this book were written and edited there.

And most important, God my Father. Any writing talent that I possess is a gift from You, and You have shepherded my life and career. May You get any and all glory for this volume.

Bold They Rise

The Feeling of Flying

On the one hand is the idea. On the other, the reality.

Sometimes the latter fails to live up to the former. The reality of expe­rience doesn’t always measure up to the way we picture it. So often in the case of space exploration, however, it is the idea that utterly fails to do jus­tice to the reality.

For example, countless descriptions of the Space Shuttle document its specifications to the smallest of details. But knowing that the vehicle stands 184 feet tall and weighs 4.5 million pounds fueled for launch doesn’t begin to capture the experience of standing at the base of the vehicle as it towers on the launchpad.

“I wasn’t intimidated by it,” recalled astronaut Mike Lounge of the first time he saw the fully stacked vehicle. “Well, that’s not exactly true. The first time we went down to the Cape on our class tour, my reaction when see­ing the pad, at seeing the orbiter and all that, is, ‘My God, this stuff’s too big. It can’t possibly fly.’ I think that’s a common reaction. I knew how big it was, but it’s different when you actually see it and you’re walking under­neath the orbiter and all this stuff. But having gotten over that, it was kind of fun to be there with the hardware. Everyone enjoys hardware over sim­ulations and paper.”

If the vehicle itself transcends expectations, NASA’s astronauts found that so, too, did the experience of actually flying aboard the Space Shuttle. Those expectations would have gradually mounted during months of mis­sion preparation and training, but the experience would truly begin in ear­nest when the highly anticipated launch day arrived.

For an astronaut, that first launch day comes only after years with NASA. Since 1978 astronauts have first been selected as “candidates” and must com­plete an initial orientation period, replete with training in almost every as­pect of the agency’s work, before becoming official members of the corps.

Then there are ground assignments supporting the program in ways that have nothing to do with getting ready for a mission.

And then, finally, years after selection, there’s the crew assignment. Fol­lowed by more training and preparation. There’s practice on the gener­al things that will occur during the mission, like launch and landing, to make sure everyone is ready. There’s practice for all the things that theoret­ically could occur during the mission but shouldn’t, the potential anoma­lies and malfunctions the astronauts have to be ready for. There’s training on mission-specific tasks, the unique things each astronaut will have to do on this particular flight. There’s preparation, working with the scientists or engineers or companies or countries responsible for the mission payloads to make sure that those, too, are ready to go. So when launch day finally arrives, it’s a long-awaited culmination of a great deal of time and effort.

Astronaut Terry Hart recalled his launch day at NASA’s Kennedy Space Center (ksc) in Florida, home of the Space Shuttle’s launch complexes: “It was a clear, cool morning there and we went through the whole morning, going through the traditions of having breakfast together, and there was al­ways a cake there for the crew before they go out. And then going into the van and realizing that all the Mercury guys went on that van, it was really a very heady experience.”

For three-time shuttle veteran David Leestma, that experience of wav­ing to people while walking out to the Astrovan, suited up and ready for launch, was a memorable moment. “We always called that the last walk on Earth,” Leestma said. “There’s always crowds of people there to see you in case you never come back or something. It was one of those little bits of kind of gruesome humor. And then you go out to the launchpad, and you’ve been through this. You’ve been there many times before, because you train in the orbiter a few times and you have countdown demonstra­tion tests and things. And this time you get to the pad and there’s nobody there. You go, ‘Ooh.’ And the vehicle is steaming and creaking and groan­ing and you go, ‘This is for real.’”

On the launchpad, the Space Shuttle is positioned vertically, its three major components having been stacked together in the enormous Vehi­cle Assembly Building at Kennedy Space Center before having been rolled out—slowly—to the launchpad atop a huge crawler. Standing tallest is the orange-brown external tank. The external tank has no engines of its own but carries the liquid fuel for the launch in two separate tanks, one con­taining liquid oxygen and the other holding liquid hydrogen. The tanks are supercooled to maintain the fuels at the cryogenic temperatures need­ed to keep them in liquid state—below minus four hundred degrees Fahr­enheit in the case of the hydrogen. Fully fueled, the external tank weighs about 1.7 million pounds.

On either side of the external tank is a slender, white solid rocket boost­er (srb), the two of which together provide the bulk of the power for the first two minutes of the launch. Once ignited, they together provide 6.2 million pounds of thrust. Their name comes from the fact that they carry their propellant—consisting largely of aluminum mixed with an oxidizer to cause it to burn—in a solid, rubbery form.

And then there’s the orbital spacecraft itself, the winged, white-and-black orbiter. Near the nose of the orbiter is the crew cabin, where the astronauts fly the vehicle and live during their mission. Farther aft is the payload bay, with its two large doors. And in the rear are the three Space Shuttle main engines, fueled by the external tank, each capable of generating a thrust of almost half a million pounds.

By launch day, the launch complex’s servicing structure has been rotated back, revealing the orbiter. The shuttle is ready for its crew. The entrance to the orbiter is through a hatch in the side of the crew cabin, near the top of the vertically stacked vehicle, almost 150 feet above the launchpad.

Leestma recalled the process of boarding the vehicle via an elevator in the launch tower and a gantry arm near the top of the structure:

As usual, people don’t say much in elevators. It’s true whether you’re in a hotel or on the launchpad. You kind of watch the numbers tick by, and instead offloors, they do everything in feet in the elevators, so you’re so many feet above sea level. And then across the gantry, and when you walk across the gantry you’re looking down into the flame trench. And you’ve been there before, but the obvious thing that’s striking you is that this is for real, we’re going to go. At least you hope we’re going to go today. . .. You get up to the White Room, the access arm, and there’s only two, maybe three people there and that’s it. There’s nobody else on the pad and everybody’s blocked off for four or five miles away. This is for real. And it’s groaning and moaning and you know that it’s going to launch, and it’s fueled and ready to go. It’s a big bomb there, sitting on the pad. And you hope that all the fire goes down and you go up, and let’s go, let’s get on it with it. It’s great…. We got strapped in, and again, the guys strapping us in were a lot of the same guys that strapped in Al Shepard on his flight [to become the first American in space during Project Mercury]. So it was a very heady time. . . . You get in and you just cant wait for it to happen.

Astronaut Jerry Ross, who was the first to launch into space seven times, said journeying out to the launchpad when the vehicle is fully fueled and ready to go is quite different than going out there any other time, not only because of the reality of the situation, but because the shuttle itself is different.

The vehicle really does give you this sense that it’s an animal that’s awake and just ready to go do something. When you go out there and the vehicle’s not fu­eled, it’s not hissing, it’s not boiling off vapors, it’s not making noises that you don’t hear, that you do hear when it’s fueled. And there’s the tremendous amount of anticipation. My first flight was the twenty-third flight of the shuttle, and I had listened to every crew come back, and I took very detailed notes of their debriefings, which were quite exhaustive early on. I listened to everything they said, and they would give us a very detailed description of what it was like, what the sensations were of launch. I put that into my databank, and I would daydream about that when I’d go running or work out at the gym or something like that. I knew it was going to be a pretty exciting ride.

The crew cabin of the shuttle has two levels. The “upper” deck is the flight deck, where the commander and pilot sit at the vehicle’s controls, with a bank of large windows in front of them. The flight deck has room for up to two more astronauts to sit during launch, and behind them are windows looking into the payload bay and the controls for the orbiter’s robotic arm.

Below the flight deck is the mid-deck, where the rest of the crew sits during launch. Once in orbit, the mid-deck serves as the primary living area for the crew, with storage lockers and the orbiter’s kitchen and bathroom and main sleeping area. The mid-deck also provides access to the vehicle’s payload bay. During launch, the mid-deck has very limited visibility, and the astronauts sitting there depend largely on word from the flight deck and the very ob­vious physical sensations of launch to know what’s going on during ascent.

Prior to launch, once the crew members have boarded the orbiter and been strapped into their seats, the waiting begins. Traditionally, the astro-

The Feeling of Flying

і. sTs-1 crew members Commander John Young (left) and Pilot Bob Crippen inside Space Shuttle
Columbia in the Orbiter Processing Facility at the Kennedy Space Center. Courtesy nasa.

 

nauts board about three hours before the scheduled launch time, lying on their backs in their chairs until launch.

Very often, this is as far as things get. Any number of issues, from unac­ceptable weather conditions to a technical glitch with the vehicle and more, can result in the launch being scrubbed and pushed back. In those cases, the astronauts are helped out of the vehicle, and work begins to prepare for the next launch attempt. “Probably one of the most frustrating things is when you get near your takeoff time, your launch time, and then you know there’s a problem, and you go, ‘Please solve it. We don’t want to wait here. Get us off the pad,’” noted Leestma. “The last people you want to have to make the real technical decision whether you go or not is the crew, because they’re always, ‘Go.’ ‘Yeah, we’ll be fine. Let’s go.’ That’s why you’ve got a whole team of folks in the launch control room doing that.”

But on other occasions, the weather does what it’s supposed to, the ve­hicle is operating properly, any number of other factors come together as they should, and launch preparations continue to proceed. Finally, as launch nears, the Space Shuttle main engines “gimbal,” or tilt, to test that they will move properly, and at five seconds before launch they are ignited to make sure all three engines are functioning properly. The vehicle contin­ues to sit on the pad, but the firing of the engines causes it to pitch slight­ly. It then rocks slightly back, a process called the “twang,” and when the stack is vertical again, at T minus zero, a spark at the top of the fuel casing of the solid rocket boosters ignites the propellant. With more than seven and a half million pounds of thrust pushing the Space Shuttle upward, it begins to move.

Shuttle pilot and commander Fred Gregory recalled the feeling of the main engines first firing, describing it as almost a nonevent. “You could hear it; you were aware of it. It sounded like some kind of an electric motor at some distance, but you looked out the window and you saw the launch tow­er there and the launch tower moved back. At least that’s what you thought, but then you realized the orbiter was moving forward and then back, and when it came back to vertical, that’s when those solids ignited and there was no doubt about it. You were going to go someplace really fast, and you just watched the tower kind of drop down below you.”

At the very beginning of the ascent, there’s the brilliant light of the en­gines, which no photograph or video can truly capture: a brightness that seems to puncture the sky. The brilliance of the flames from the engine is dramatic during the day, and far more so when they light up the sky at night. Payload specialist astronaut Charlie Walker recalled the experience of launching on the Space Shuttle in the dark:

At night, you look outside, and this launchpad is a blue gray from the xenon light reflections bouncing off of it, with a completely black background behind it. All of a sudden the launchpad brightens up with the solid rockets igniting. The launchpad brightens up to a yellow gray, but then the whole background, suddenly there’s like a sunrise that’s happened over Florida. You can see the Flor­ida landscape for miles back that way. Sure, the sky is still black, but suddenly Florida has been illuminated by a new sunrise. I can see the Florida country­side, and it’s a yellow, white-yellow-orange color, the coloration of the brilliant, hot flame from the solid rocket boosters.

Like Gregory, Jerry Ross recalled that, while he was aware when the main engines first ignited, things didn’t really get exciting until the solid rocket boosters fired.

As the shuttle’s main engines came up, you could really feel the vibrations starting in the orbiter, but when the solid rocket motors hit, when they ignite, it’s somebody taking a baseball bat and swinging it pretty smartly and hitting the back of your seat, because it’s a real “bam!" And the vibration and noise is pretty impressive. The acceleration level is not that high at that point, but there is that tremendous jolt as the solid rocket motors ignite, and you’re off I’ll never forget the vibrations of the solid rocket motors. As we accelerated in the first thirty seconds or so, the wind noise on the outside ofthe vehicle just became really intense, like it was just scream­ing. It was screeching on the outside. I was already thinking about “what am I do­ing here" before then, but [it was] just a sheer, incredible experience ofthe energy.

In many ways the flight deck, with its large windows, is the superior seat­ing for experiencing the launch. In one way, however, the mid-deck has the advantage. Since the pilot and commander are busy with the tasks of mak­ing sure the vehicle is operating properly during ascent, they don’t have the luxury of stopping to really take in the experience of the launch. While the astronauts on the mid-deck don’t have the same view as those on the flight deck, they have the freedom to focus more on the sensations. Hart, for exam­ple, recalled being able, as a mission specialist, to really enjoy the experience.

You talk a lot [to other astronauts about what launch is like], obviously, and you see a lot of pictures, and you think about it a lot, so you think you’re pret­ty well prepared and you probably wont have too many surprises, but I had a couple of surprises. The shake, rattle, and roll for the first two minutes, that was about what I thought, maybe even a little bit less than what I thought it would be, because the solid rockets kind of have a “whoof-whoof" [rumble]. You don’t really hear it; you more feel it. It’s like a very low-frequency rumble, and just a tremendous sense of power as you lift off and all.

Another part of the experience that simply cannot be replicated on the ground is the pressure of the g-forces during ascent, according to sts-6 commander P. J. Weitz: “The value of our simulators ends when those en­gines light and you lift off. They try to fake you out a little bit by tipping the Shuttle Orbiter Simulator and that, but it doesn’t compare with three shuttle main engines and two solids going. As I tell people, I said, ‘You know you’re on your way and you’re going somewhere and you hope they keep pointed in the right direction, because it’s an awesome feeling.’”

Weitz compared the launch of the Space Shuttle to the launch of a Sat­urn IB, which he took into space on the first Skylab mission. The Saturn, he said, produced about half again as much acceleration force as the shut­tle’s three gs, and the force was felt in somewhat different ways on the two vehicles. In the Saturn, the thrust was “actual,” or directly in line with the vehicle, so the crew was pressed directly back into the couches. With the shuttle, on the other hand, because of the way the orbiter is stacked on the external tank, the thrust from the main engines is offset from the ve­hicle’s center of gravity, meaning that the crew members aboard felt the pressure pushing them not only into the back but also into the bottom of their seats.

After clearing the launchpad, the shuttle begins to roll so that the orbit – er is below the external tank, to better allow its engines to offset the tank’s weight. Around one minute into flight, the shuttle encounters “Max Q,” the period in which the increasing velocity of the vehicle produces the max­imum amount of pressure on the shuttle before the decreasing resistance of the atmosphere reduces that pressure. To reduce the strains of the pressure of Max Q, the vehicle throttles down its engines and then, seconds later, past the point of maximum pressure, throttles back up.

Just over two minutes into the launch, the solid rocket boosters separate from the vehicle, and the orbiter and external tank continue toward orbit. The solids deploy parachutes and land in the ocean, where recovery ships locate them and bring them back for refurbishment and reuse.

“At the solid rocket motor separation. . . there was this brilliant orange flash, orangeish-yellow flash across the windscreen, and then the solid rocket motors are gone,” Ross recalled. “As the solid rocket motors tailed off, like at a minute forty-five or so, it almost felt like you had stopped accelerating, almost like you’d stopped going up. At that point we were already Mach 3-plus and well above most of the sensible atmosphere at that point, some twenty miles high or so. And at solid rocket motor jettison, then you’re at four times the speed of sound and twenty, twenty-five miles high.”

Hart also recalled the separation of the solid rockets as a memorable ex­perience. For the first two minutes of ascent, the g-forces that the crew ex­periences have been building up, and then, at srb separation, they drop off dramatically.

Very quickly, then, the solid rockets taper off and separate, and that was the first surprise I had. . .. The sensation that you have at that point I wasn’t quite pre­pared for, because you go from two and a half gs back to about one and a half. Well, when you get used to two and a half, and it feels pretty good. You’re going somewhere, you know. When you go back to one and a half, [it] feels like about a half. So you don’t think like you’re accelerating as much as you should be to get going. And, of course, I had worked the main engine program anyway, so I was very familiar with what the engines could do or not do. And I think in the next minute, every five seconds I checked the main engines to make sure they were running, because I swear we only had two working, because it just didn’t feel like we had enough thrust to make it to orbit. But then gradually the ex­ternal tank gets lighter, and as it does, of course, then, with the same thrust on engines, you begin to accelerate faster and faster. So after a couple of minutes I felt like, yes, I guess they’re all working.

Ross also had the experience of worrying that all main engines were not working when they actually were.

I literally had to look to see that the three main engines were still working, be­cause it became so smooth, and it almost felt like you weren’t going anywhere;

you weren’t accelerating at all. . . . At one point I can remember looking back behind me out the overhead windows again. In artists’ renditions of the flames coming out of the three main engines, it’s a nice, uniform cone of fire back there and stuff. Not true. The fire was all over the place. It was not static. It was dancing. It was not uniform. And again you go, “Is this thing working okay?" You don’t know what to expect.

As the shuttle nears the end of its powered ascent, with the bulk of the atmospheric drag behind it, it begins to accelerate dramatically. “As we got up to about the seven-and-a-half-minute point, then, is when you get to the three gs of acceleration, and that’s a significant acceleration,” Ross said.

It feels like there’s somebody heavy sitting on your chest, and it makes it pret­ty hard to breathe. I mean, you kind of have to grunt to talk, and you’re just waiting for this three gs to go away. . . . You’re accelerating at 100 feet per sec­ond, which is basically like going from 0 to 70 miles per hour every second. So it’s pretty good. And then at the time that the computers sense the proper condi­tions, the main engines. . . shut off and you’re in zero g. And for me, the first flight, sitting in the back seat, I had the sensation of tumbling head over heels, a weird sensation. And it was the three-g transition, from three gs to zero gs…. But as soon as I got out of the seat, then I was okay.

The main engine cutoff, or meco, comes around eight and a half minutes into the launch, and shortly thereafter the external tank sepa­rates from the orbiter and reenters Earth’s atmosphere. As the only ma­jor component of the shuttle stack that isn’t reusable, the external tank burns up on reentry.

Gregory explained how he felt in that moment, when the main engines cut off and he was floating in the microgravity of space: “The first indica­tion that this was not a simulation was when the main engines cut off and we went to zero g, and though [Steven] Hawley, I think, had been attrib­uted with this comment, it was a common comment: ‘Is this space? Is this it? Is this real?’ And it was an amazing feeling. I’d never sensed anything like this before. So this sensation of zero g was like a moment on a roller coaster, when you go over the top and everything just floats.”

Hart described being surprised once in orbit, but unlike Ross and Greg­ory, not by the experience of zero g.

The zero g I was pretty well prepared for. As a fighter pilot and the experience at NASA in the zero-g trainer, you’re pretty familiar with what it feels like to be weightless. But what I wasn’t prepared for was the first look out the window. You don’t know what black is until you see space. I mean, I was startled with just how black it was. You don’t see stars. You could barely see the moon; it’s be­cause there’s so much light coming off the Earth and off the tiles of the shuttle, that there’s a tremendous ambient light from all those sources, so your eyes are constricted greatly. And then because of that constriction, when you look into space you can’t see the stars or anything. I mean, it’s like really black. It’s pal­pable. You think you can almost reach out and touch it. I don’t know quite how to describe it. It’s sort of like black velvet, but it’s just totally palpable. . . . I guess I knew that I wouldn’t be able to see the stars when we were on the day side of the Earth. But still, when you look out there and see the blackness, it re­ally was striking to me.

While most astronauts report experiencing an overwhelming excitement or elation upon their first arrival in orbit, Fred Gregory jokingly recalled an odd bit of disappointment stemming from his first ascent. “Since we had trained constantly for failures, I anticipated failures and was somewhat dis­appointed that there were no failures, because I knew that any failure that occurred, I could handle. It was where I slipped back into an ego thing. I anticipated failures that I would correct and then the newspaper would say, ‘Gregory Saves Shuttle,’ but heck, none of that happened. It just went up­hill, just as sweet as advertised.”

As pilot, Gregory said his main job once the vehicle was on orbit was to make sure it was working properly. Since there were no major issues, he found that he had frequent opportunities for looking out the window. Said Gregory,

You immediately realize that you are either a dirt person or a space person. I ended up being a space person, looking out in space. It was a high-inclination orbit, so we went very low in the southern hemisphere, and I saw a lot of star formations that I had only heard about before and never seen before. I also saw aurora australis, which is the southern lights. I was absolutely fascinated by that. But if you were an Earth person, or dirt person, you were amazed at how quickly you crossed the ground; how, with great regularity, every forty-five min – utesyou’d either have daylight or dark; how quickly that occurred, about seven miles per second; how quickly you crossed the Atlantic Ocean.

Although he was a self-described “space person,” Gregory still enjoyed occa­sionally gazing down on Earth below him and found it a fascinating experience.

The sensation that I got initially was that from space you cant see discernible borders and you begin to question why people don’t like each other, because it looked like just one big neighborhood down there. The longer I was there, the greater my “a citizen of" changed. The first couple of days DC was where I con­centrated all my views, and I was a citizen of Washington DC. I was confused because I thought everybody loved DC, but [Bob] Overmyer was from Cleveland [Ohio], and Don Lind was Salt Lake [City, Utah], and Norm [Thagard] was Jacksonville, Florida, and Lodewijk [van den Berg] was the Netherlands, and Taylor Wang was Shanghai [China], so each had their own little location for the first couple of days. After two days, I was from America, looked at Ameri­ca as our home. Taylor, China. Europe for Lodewijk. And after five or six days, the whole world became our home.

During his flight, Gregory developed a sense not only of Earth as a whole being his home but of just how interconnected the global community truly

is, and the extent to which all people are sharing one planet.

You could see this kind of sense of ownership and awareness. We had noticed with interest the fires in Brazil and South Africa and the pollution that came from Eastern Europe, but it was only with interest. After five or six days, then it was of concern, because you could see how the particulates from the smokestacks in Eastern Europe, how that circled the Earth and how this localized activity had a great effect. When you looked down at South Africa and South America, you became very sensitized to deforestation and what the results of it was with the runoff how it affected the ecology. Then you’d have to back up and say, well, this is not an intentional thing to destroy; this is something that they use coke as part of their process, and in order to get coke, you’ve got to burn. So you be­gan to look at things from different points of view, and it was a fascinating ex­perience. So that was the science that I was engaged in, but never anticipated

it. And it was a discovery for me, so as each of these other great scientists who were with us discovered something that they had never anticipated, I also did, and I think the whole crew had.

In order to live and work in space during their missions, astronauts must learn to adapt to the microgravity environment, and that adaptation varies from individual to individual. Part of the adaptation is simply learning to get around; moving through the vehicle without gravity is an entirely dif­ferent process than walking through it on the ground. For many other as­tronauts, adaptation involves a physical unease as the vestibular system ad­justs to the lack of the orienting influence of gravity.

While it may take different amounts of time for astronauts to be back to 100 percent, most are at least functioning fairly quickly, Gregory said.

Whatever the adaptation was, within a day, everybody had adapted to it and so it was just a matter of working on all the programs and projects of the proj­ects that you had. The body very quickly adapted to this new environment, and it began to change. You could sense it when you were on orbit. You learned that your physical attitude in relation to things that looked familiar to you, like walls and floors and things like that, didn’t count anymore, and you translat­ed [from thinking about] floors and ceilings and walls to [thinking]your head is always up and your feet are always down. That was a subconscious change in your response; it was an adjustment that occurred up there. You also learned that you didn’t go fast, that you could get from one place to the other quickly, but you didn’t have to do it in a speedy way. You always knew that when you started, you had to have a destination, and you had to have something that you could grab onto when you got there. But, again, this was a transition that occurred, perhaps subtly, but over a very short period of time. I can remember we all kind of joked up there that we had become space things, and we were no longer Earth things anymore. The first couple of days, a lot of bloated faces, because there was no gravity settling of the liquids. But after a couple of days, you lost that liquid in your body, and you looked quite normal. So it was a fas­cinating experience. I think it was surprising to us how quickly we adapted to this microgravity environment.

With launch complete and their bodies adapting to space, the astronauts would go about their mission, spending days on any variety of different tasks carried out by shuttle crews during the early years of the program. Finally, though, the time would come to return to Earth. The orbiter would turn backward relative to its velocity and fire its engines to slow itself down, be­fore rotating back to begin its descent.

The Feeling of Flying

2. During sts-8, Commander Richard “Dick” Truly and Mission Specialist Guion Bluford sleep on
Challengers mid-deck. Courtesy nasa.

 

The experiences of launch and landing are very different, Gregory said. Ascent is relatively quick and marked by rapid changes in the g-forces ex­perienced by the crew. Landing, on the other hand, is far more gradual.

On reentry, it is entirely different. Though it takes eight and a half minutes to get up to orbit, it takes more than an hour to reenter, and it feels very similar to an airplane ride that most people have been on. You get an excellent view of the Earth. If it’s night when you reenter the atmosphere, then you see a kind of a rolling plasma over the windows. . . . But other than the onset of g that oc­curs at less than 400,000 feet above the Earth, it is like flying in an airplane. The sensations that you have are very similar to a normal domestic airplane flight. You’re going pretty fast, but you are not aware ofit because you’re so high.

It’s an amazing vehicle, because you always know where you are in altitude and distance from your runway. You know you have a certain amount of en­ergy and velocity, and so you also know what velocity you’re supposed to land, and you can watch this amazing electric vehicle calculate and then compen­sate and adjust as necessary to put you in a good position to land. We normal­ly allow the automatic system to execute all the maneuvers for ascent and for reentry, but as we proceed through Mach 1, slowing down for landing, it is customary for the pilot, the commander, to take command of the orbiter and actually fly it in, using the typical airplane controls. But, you know, as I look at it, the ascent is very dramatic. It’s very fast, a lot of movement, but quick. The entry is more civilized but exposes the orbiter to actually a greater danger than the ascent, as far as the influence of the atmosphere on the orbiter. The temperatures on the outside of the orbiter really get hot on reentry, and that’s not the case on ascent.

Astronaut Charlie Bolden flew on the Space Shuttle four times, two of those as commander, and became nasa administrator in 2009. A former naval aviator, Bolden described landing the Space Shuttle as a unique experience. “The entry and landing is unlike almost anything you ever experience in any other kind of aerospace machine because it’s relatively gentle,” Bolden said.

In terms of g-forces and stuff like that, it’s very docile. Unless you do something wrong, you don’t even get up to two gs during the reentry, the entire time of the reentry. When you bank to land, you come overhead the landing site, and then you bank the vehicle and you just come down like a corkscrew. . . . It feels like you’ve got gorillas sitting on your shoulder because you’ve been weightless for x number of days. And so it’s just a really different feeling. You have to hold your head up because you’ve got this big old heavy helmet on and it probably weighs [a few]pounds, but it feels like it weighs a hundred. It takes a little bit ofenergy to get your hands up offthe console, because once you start feeling gravity again, your hands just kind ofgo down and they want to stay there; everything does. So the two pilots on board are doing a lot of isometric exercises all the way down.

Even when an astronaut lands the shuttle for the first time after a mis­sion, Bolden said, it already feels very familiar because of all the training in preparation for the missions. “It’s like you’ve done it all your life, because you have,” Bolden said.

You’ve done it thousands of time by now in the shuttle training aircraft for real, and you’ve done it probably tens of thousands of times in the simulator. So it doesn’t look abnormal at all; it’s just something that you ’re accustomed to. When you touch down, if you do it right, again, you hardly know you touched down. As big as the orbiter is, the way that we land it is we just get it into an extreme­ly shallow approach to the landing, and so it just kind of rolls out on the run­way, and if you do it right, you all of a sudden notice that things are starting to slow down real quick and you’re hearing this rumble because the vehicle’s rolling down the runway on this grooved runway. So you know you’re down, put the nose down and step on the brakes and stop. That’s it. And then you go, “Holy G. I wish it hadn’t been over so quick. " I don’t think it makes a difference how long or how short you’ve been there, it’s over too quick. You’re ready to come home, but once you get back, you say, “Boy, I wish I had had a few more days," or something like that. And for me, my last two, being the commander and actu­ally being the guy that had the opportunity to fly it to touchdown, was thrilling.

Once the landing is completed and the orbiter is safely back on Earth, the crew begins the process of reacclimating to the planet’s strong gravity after days of feeling weightless. Charlie Walker, the first commercial pay­load specialist, who flew on the shuttle three times, recalled waiting in the orbiter at the end of the mission.

The guys on the flight deck were going through the closeout procedures. Ground crews were closing in. We sat unstrapped, but we would sit in our seats for anoth­er ten, fifteen minutes as the ramp was brought up, the sniffers checked for am­monia leaks and/or hypergolic propellant leaks, found none, and put the stairway [up to the hatch], and opened the hatch. All that time, all of us are beginning to get our land legs back, unbuckle, start to try to stand up. “Ah, this doesn’t feel good yet. Wait a little bit longer." So you kind of move around, move your arms first, your feet first, your legs first, then stand up, make sure you’ve got your bal­ance back. The balance is the one thing that you just don’t have. Again, the brain hasn’t been utilizing the inner ear or senses of where the pressure is on the bottom of feet, for instance, to use as cues to balance itself against gravity. It hasn’t done that for a week. So you’ve got to carefully start through all that and consciously think about balance and consciously think about standing up, and we very con­sciously do that, because the last thing you want to do, in front ofhundreds ofmil – lions of people watching on television, is to fall down the ramp leaving the orbiter.

Normally, on Earth, the body works hard to make sure the brain is ad­equately provided with blood. From a circulatory perspective, the brain, the part of the body that most needs blood, is located inconveniently at the top of the body, so the heart has to pump blood against gravity to get it there. In orbit, on the other hand, blood flows much more easily to the head, but it doesn’t fill the legs the same way without gravity pulling blood into them. Astronauts develop bloated heads and “chicken legs” due to the body’s confusion over how to distribute blood without gravity. The body takes the increased fluid flow to the head as a sign that it is overly hydrat­ed and begins to shed what it sees as excess fluid. After the return to Earth, fluid redistributes again, which can cause problems.

“The body adapts by, among other things, letting go of a lot of fluid, about a liter of liquid, which makes you clinically dehydrated while you’re in space, except the whole condition of the body is different up there, so you’re really not dehydrated in that environment,” Walker said.

But if you come back without replacing that liter of fluid, then you are dehy­drated. You try to stand up with not so much fluid to go to the head, and so you literally could pass out. Nobody did that, but I know I had sensations of light­headedness for the first few minutes until I just literally worked at getting my balance back and focusing attention, and the body was adapting all that time, too. But leaving the spacecraft, I was holding onto the handrail as I went down the stairway. Got to the bottom of the stairs, and I was walking like a duck, because I was trying to keep my balance.

Once they’ve adjusted enough to walk, crew members board the Astrovan, which takes them to the medical quarters for postflight medical exams and a shower. Walker said it felt good to take a shower after days without one.

Every sensation for the next many hours, normal sensations of water running over you in a shower, [felt] strange. Because again, here this water’s hittingyou, and it’s running down. And hours later, I found that I still could at any mo­ment just think about the sensations in my body, and it was odd to feel this pull down toward the surface of the Earth, to be stuck to the surface of the Earth. [When I flew], it was still fairly new to hear comics or some wag note that this or that “sucks." [Coming back] the astronauts were saying, “Well, the Earth re­ally does suck." So it keeps me drawn right down to the surface. Gravity is real­ly real, and it stands out in your mind to, again, the freedom of weightlessness when you’ve had that opportunity. And that was just very much on my mind. I remember even a day, two days later, probably like a day later at a meal, I was sitting down, and I could not easily figure out whether I should sit back against the back of the seat or lean forward, because my head was telling me I was leaning forward at an angle, and, in fact, I was sitting almost straight up and down. So the inner ear is still adapting to its own senses and the body’s cues to orient itself and still doesn’t have itself figured out completely yet.

Even if an astronaut spends only a few days in orbit after a lifetime of living in the gravity of Earth, habits developed during those few days of weightlessness can persist for a little while after the mission. “I also remem­ber waking up the next morning back here in Houston, waking up and go­ing into the bathroom and wanting to brush my teeth, and I did that, and I remember letting go of the toothbrush, and it fell to the sink top, and I probably laughed,” Walker noted.

Then I pick up the cup of water to rinse my mouth out, and then proceed to let that cup go again. It’s like, again, you’re still thinking weightlessness, and you’re really used to that. Finding the situation where gravity is ever-present is just such an interesting experience, because now, again, you’ve had that contrast of a dif­ferent place where that wasn’t part of the environment and you note when you get back how remarkable and how constraining gravity is. . . . We’ve all grown up for some decades, before we go fly in space, in gravity, and it’s just natural. Except it is programmed in, and that programming is submerged with new habits that you gained to work in weightlessness, and you have to pull that pro­gramming back, or the brain does, and it does so at different rates, I think. So within tens of minutes, you can walk comfortably. You may look a little odd, because you’re not walking as expertly as you had done for twenty, thirty, forty years before. It takes a few more hours, maybe a couple hours to do that. But you can walk, so balance comes back pretty darn quickly. But it’s probably the nonautomatic stuff like I’ve remarked about just automatically leaving a glass hanging in the air, thinking it’s going to stay there. You just get into habits there that are semiconscious, and it takes a little while for the body and the brain to let go of that and to relearn that, no, I’m stuck here again to the surface of the Earth. I’ve got to put the glass right up here on the table directly.

In the Beginning

Arguably, it could only have happened when it did.

Astronaut John Young, who would go on to become the commander of the first Space Shuttle flight, was standing on the surface of the moon dur­ing the Apollo 16 mission in April 1972 when he heard the news that Con­gress had approved vital funding for the development of the shuttle in its budget for fiscal year 1973. He reportedly jumped three feet into the air on the lunar surface upon hearing the news.

The Space Shuttle would be the most complex piece of machinery built by humankind. It was an incredible challenge and a daunting undertaking. At another point in history, a decade earlier or even a decade later, it might have seemed too challenging, too ambitious. But the project was born when men were walking on the moon. From that perspective, anything was possible.

It would be, far and away, the most versatile spacecraft ever built. But to many of the early astronauts who were involved in its creation, it was some­thing even more fascinating—an aircraft like no other. Talk to the astro­nauts brought in as pilots during the 1960s, and there’s a fair chance they’ll refer to the orbiter as “the airplane.” Many of them will talk about its de­velopment not in terms of rocket engines and life-support systems but in terms of avionics and flight control systems. They had been pilots, many of them test pilots, and they had come to NASA to help the agency fly capsules through space. But now—now they were aircraft test pilots again, helping to design an aircraft that flew far higher and far faster than any aircraft before.

Since the selection of the first astronauts, members of the corps had been involved in the development of new spacecraft and equipment, providing an operator’s perspective. These were the people who would have to use the things that the engineers were designing, so it was their job to give the en­gineers feedback on whether the things they were designing were actually usable. For much of the time the Space Shuttle was being developed, most

of the astronaut corps was grounded, with only a dozen flying between the last moon landing in 1972 and the first shuttle flight in 1981. As a result, there was plenty of opportunity for astronauts to be involved in the devel­opment of the shuttle, and they participated more in the development of this vehicle than any before.

Even so, there were some at nasa with the idea that the moon would be just the first step into the solar system, who were concerned about what the shuttle wouldn’t be able to do—go beyond Earth’s veritable backyard.

In January 1973 astronaut T. K. Mattingly was assigned to be head of As­tronaut Office support to the shuttle program. This was around the same time that the contracts were being awarded to the companies that would be responsible for making the shuttle’s various components. Mattingly, who had orbited the moon on Apollo 16 while Young was walking on it, recalls talking to Deke Slayton, the head of flight crew operations at nasa’s John­son Space Center (jsc) in Houston, Texas, about the assignment. “When I got back from Apollo 16, Deke asked me, he said, ‘You know, there’s only one more flight, so if you really want to fly again anytime near-term, you might want to take the backup assignment on [Apollo] 17,’ he said. ‘Chanc­es aren’t very good, but we do know that we replace people occasionally. So if you would like to have that chance, you can do it, or you could work on the shuttle program.’ Really, I hadn’t paid much attention to it,” Mattingly said of the shuttle program at that point.

I kind of knew the work was going on, but I didn’t know what it was, because my ambition had always been—I didn’t think I would go to [walk on] the moon, but I was really hoping that I’dget to be on the Mars mission, which I was sure was going to happen the following year. To a young kid, it just seemed obvious that the next step is you go to the moon, then you sharpen your tools and you go to Mars, and I thought, “Boy, that’s where I’d like to go. ”

Even by then it was becoming obvious that that wasn’t really a likely propo­sition. I wasn’t enthused about the shuttle because I still thought going to Mars was the next step. I believe that we needed to build a space station first so we could have hardware, which would gather years oflifetime experience while we could get to it and fix it, and we could build the transportation system while we’re gaining the experience with a space station. All of that architecture was obviously politically driven, and they were having to fit into a tighter budget.

There really was not a great swell of emotion or enthusiasm for things follow­ing Apollo in the political arena, nor in the public arena, for that matter. So I think they had to walk some very, very tight lines in order to keep the program going, and so they chose the Space Transportation [System] as the way to go.

George Mueller, the head of manned spaceflight at NASA during the Apollo program and the man many recognize as the father of both the Sky – lab space station and the Space Shuttle program, said that, even with the development of the shuttle, human exploration of other worlds remained the ultimate goal. “It became clear that the cost of getting into orbit was the driver for all future programs. I began to think about, how do you get the cost down. In air travel, you can’t fly from here to London and then throw the plane away when you get to London. What we came up with was a completely reusable vehicle. We had every intention of going back to the moon. What we were doing was going into low Earth orbit and estab­lishing a base there; it was a requirement for reaching our long-term goal.”

Former Johnson Space Center director Chris Kraft recalled the approval of the shuttle as “a real come-down for NASA.”

We, the powers that be at NASA, had grand visions of going back to the moon, having bases on the moon, and on to Mars. They made very significant reports on what the future of NASA could and should be. But when the Nixon admin­istration decided that the limitations of the budget in his [thepresident’s] mind would not allow us to do those kinds of grand things in space, that’s when the powers that be in NASA decided, well, what is the one thing that we need to start the next generation of spaceflight? And that is we need a cost-effective launch system. That’s the first thing we need. If we’re going to go into orbit and do grand things, or if we’re going to put things in orbit and rendezvous and go other places, what we need is a good truck. We called it a truck, at times. And so that’s how we arrived at that being the next step in the space program being a reusable, therefore fly-back vehicle. We signed a fixed-price, seven-and-a-half – billion-dollar contract to build the Space Shuttle, and that was to be provided with annual increases in the budget for inflation. We never got the first piece of inflation at any time in the history of the budget of the shuttle. They welshed on that guarantee immediately, and furthermore, they delayed the program a year and did not give us any relief on the total cost, on the total fixed cost. They didn’t want the money in the budget that year, just that simple. So in the his­tory of the shuttle program, up until we made the first flight, we were always pushing a bow wave of being behind budget.

Many in the astronaut corps had doubts as to what the shuttle decision would mean for the future of exploration. Mattingly considered leaving nasa completely, believing he would probably never leave Earth orbit again.

I went up to pay courtesy calls to the navy after we got back, and John War­ner was then secretary of the navy, and we made a courtesy call to him. He was all enthusiastic. He says, “You navy [astronaut] guys need to come back, and we’ll give you any job you want. You pick it. Whatever you’d like. You want a squadron? You want to do this? Just tell me. It’s yours. ” Boy, my eyes lit up, and I thought, “Wow. ” One of my escort officers was a captain in the Penta­gon. He went back and told his boss, who was the chief of naval aviation, what Warner had said, and very quickly I had an introduction to the chief of na­val aviation, who made sure that I understood that despite what the secretary had said, in the environment we were in, I was not going to come in and take over his squadron. He’d find a place for me, he’d give me a useful job, but don’t think that with the Vietnam War going on and people earning their positions the hard way, that I was going to walk in there and do that. He says, “The sec­retary means well, but we run the show. ”

So armed with that piece of information that if I went back on real navy duty at that point I was probably not going to find a particularly rewarding job, I thought the opportunity to get in on the shuttle at the beginning and go use some of the experience we gained would be useful, so I told my sponsor I’d do whatever the navy preferred I do. After all, they gave me my education and everything else that mattered. “So you tell me, but if I had a vote, I would say why don’t I stay because the shuttle program’s only going to take four years. ” That’s what we were advertising. You know, four years, that’s not all that long. So after a significant amount of discussion within the navy side ofthe Pentagon, they said, “Okay. Well, we agree. You probably can contribute more if you stay there. ” So that lead me to stay with the shuttle program, and so the beginning of that was a period of a great deal ofthe turmoil of getting started.

Step one of designing a Space Shuttle was deciding exactly what a Space Shuttle should be designed to do. Its official name, the Space Transporta­tion System, summarized a basic part of the requirement. The shuttle would transport astronauts and cargo from the surface of Earth into space and back. It also was to be, as much as possible, reusable. The idea was that cre­ating a spacecraft that was as reusable as possible would cut down on what had to be built for each launch, and thus on the cost of each launch. Low­er the cost of putting a pound of material in orbit, and you can put more pounds of material in orbit. The space frontier opens up.

“We had a general idea of what specifications the shuttle was supposed to be, but in those days it was substantially larger and more aggressive than what we know today,” Mattingly said. “So we went through this require­ments refinement where everybody broke up into groups to go lay out what they had to do, and it evolved into something we called design reference missions. Rigidly, the idea was, we knew the shuttle was going to last for decades, and we knew nobody was smart enough to define what those mis­sions that would come after we started were going to evolve into. So we took great pride in trying to define the most stressful missions that we could.” Mattingly said the program initially outlined three types of possible mis­sions. One was for the shuttle to be used as a laboratory. “We laid out all the requirements we could think of for a laboratory—the support and what the people need to work in it, and all that kind of stuff,” Mattingly recalled. A second type of mission was defined as deploying a payload on orbit. “That was to be one that launched and had the manipulator arm and cradles and all of the things necessary to do that.”

Then there was the idea of a polar mission. Such a mission would involve putting the shuttle in a polar orbit—leaving the launch site and heading into a north-south inclination that would cause it to orbit from one pole to the other. A satellite in polar orbit would be able to fly over any point on the surface of Earth—a valuable capability for intelligence gathering. “The polar mission was really shaped after a DoD [U. S. Department of Defense] requirement,” Mattingly said.

The original mission, as I recall, was a one-rev mission. [A “rev" is essentially one orbit around Earth.] You launched, got in orbit, opened the payload bay doors, deployed a satellite, rendezvoused with an existing satellite, retrieved it, closed the doors, and landed. And this was all going to be done in one rev or maybe it was two revs, but it was going to be done so that by the time anyone knew we

In the Beginning

were there, it was all over. Well, we worked on that mission and worked on it and worked on it, and finally it became [two different design reference missions]. We just couldn’t figure out how to do it all on one short timeline.

The military design reference missions were a response to a political exigen­cy NASA had learned to deal with during the 1970s. Most notably, in develop­ing the Skylab space station, nasa found itself competing for funding against the air force, which was seeking money at the same time for its Manned Or­biting Laboratory program. Although the two programs were very different in their goals, they shared enough superficial similarities that Congress ques­tioned why both were necessary. With the shuttle, nasa hoped to avoid a re­peat of this sort of competition, and have an easier sell to Congress, by gain­ing buy-in for the idea from the military. According to astronaut Joe Allen,

Leadership in the early 1970s decided the only way the Apollo-victorious NASA would be given permission to build a reusable space transportation system is that there be identified other users for the system other than just the scientists. This na­tions leadership identified the other users as the military. The Space Shuttle would be used to carry military payloads. The military has its responsibilities, and they said, “All right. If our payloads are going to go aboard, we do have one require­ment; that is that your Space Shuttle be able to take the payloads to orbit, put them there, and land back at the launch site after making only one orbit of the Earth. ”

The need for quick, polar missions greatly affected the design of the shut­tle, yet interestingly the Space Shuttle never flew a polar-orbit mission. “At face value, that doesn’t seem all that difficult to do,” Mattingly said of the polar-orbit missions,

but what it meant was, the shape of the orbiter went from being a very simple lifting body-type shape, with very, very small wings, to a much larger vehicle with delta-shaped wings. I don’t know the exact numbers, but the wings that go to orbit and come home again [make up a large portion of] the weight of the vehicle, and they’re never fully used; only the outermost wingtips are used. All that vast expanse—with all that tile, and all the carbon-carbon [carbon-fiber – reinforced carbon] along the leading edge—is never used. It would be used if it were to go to space in a polar orbit and then come home. It would be used to gain the fifteen hundred miles of cross-range that one needs because the Earth moves fifteen hundred miles in its rotation during the time you’ve gone once around. So you have to have some soaring ability. That’s what these large wings are for. The Space Shuttle would have cost much less money. It would cost much less to refurbish each time. Still, it would not be an economic wonder, but it would be economically okay, were it not for these huge wings. Of course, that requirement, in hindsight, was never used, was never needed, but the current Space Shuttle will forever be burdened with these wings.

Mattingly also said that the design missions established the capabilities that the Space Shuttle system would need to have. Each specification let to a variety of trickle-down requirements, and gradually the vehicle began taking shape.

These requirements we set really had some interesting things. Some of them were politically defined, like you’ll land at any ten-thousand-foot runway in the world. That’s all it takes. In selling the program, they had to appeal to just every constituency you could find to cobble together a consortium of backers that would keep the program sold in Congress. People don’t recognize how that rip­ples back through a design into what you really get, and, of course, by the time you know what you’ve got, the people who put those requirements in, they’re history. So it’s interesting. But that ten-thousand-foot runway requirement set a lot of limits on aerodynamics and putting wings on the airplane. The cross­

range—that was the airforce requirement for this once-aroundpolar mission abort—that sized the wings and thermal conditions. That precluded us from using a design called a lifting body that the folks out at Edwards [Air Force Base, California] had been playing with and had demonstrated in flights. It was structurally a much nicer design, but you just couldn’t handle the aerody­namic characteristics that were required to meet these things. So we had a ver­tical fin on this thing and big wings, and it’s a significant portion of shuttle’s weight, and the maintenance that goes with it is attributed to the same thing.

Mattingly had the unique vantage point of watching the shuttle program evolve from a concept through logistical support into its mature state, he recalled. “I look back and I say, ‘Well, we know what we started to do, and we know what we have, and they’re not always the same. Why?’ Because it was an extraordinary job. Apollo was a challenge because it was just so big and it was audacious, and time frame was tight, and all of those things.” But in many ways, Mattingly said, the shuttle was even more challenging.

Essentially, it was so demanding that all of the engineering and ops [operations] people. . . generally stayed on. We didn’t have a lot of technical attrition after Apollo. At least that’s my impression. At least the middle-level guys all stayed, and they kept working it because they recognized that the shuttle was a far more challenging job than Apollo in many technical senses.

The part ofthe shuttle that was different was Apollo was a collection of boxes. If you had a computer, you could build it, you could test it, you could set it out and do it all by itself. You had a second stage. You could build and test the whole thing by itself. Well, with the concept of this reusability and integration, you didn’t have anything until you had everything. There was no partial thing. There was nothing that was standalone. I remember we were trying to buy off-the-shelf tacans [Tac­tical Air Control and Navigation systems], an airplane navigational system, and as part of this integration process, rather than take the tacan signal that an airplane generated in those days and used for navigation, we stripped it all out and put in all our own software so that this off-the-shelf tacan box was absolutely unique. There was nothing else. And it was part ofthe philosophy of how we built this system.

Despite the areas where the shuttle fell short of the original requirement – based specifications, Mattingly said NASA ended up with a very robust and versatile vehicle because of how ambitious the original discussions were. “At

In the Beginning

4- Possible configurations considered for the Space Shuttle, as of 1970. Courtesy nasa.

the time we were doing this and putting all these requirements on there, we were actually, I think, quite proud of having had the foresight to look at all of these things. Today you can hardly think of a mission. . . you’d like it to do that it can’t do. It is an absolutely extraordinary engineering piece, just unbelievable. The shuttle really did fulfill almost all of the requirements that we were tasked to put into it.”

The shuttle went through a variety of widely different configurations during its early development. An inline version would have had the orbiter on top of a more traditional rocket booster, which would use parachute recovery to make it reusable. Another version would have had the orbiter launched atop essentially another space plane that would fly back to a ground landing site.

Discussions were held as to whether the primary fuel tank, which ended up being the external tank, should be inside the orbiter or not. There were trade-offs, according to Chris Kraft, the Johnson Space Center director at the time. Putting the tank inside the orbiter would have required that the orbiter be much larger but would have greatly increased the reusability of the shuttle system. However, Kraft said, the ultimate limitation was the dif­ficulty of designing an integrated vehicle that wouldn’t suffer substantial damage to the fuel tank during landing.

Another major issue that had to be figured out early on was what sort of escape system should be provided for the crew. The Mercury and Apol­lo capsules both had powerful solid rocket motors in the escape towers at the top of the vehicle that would have been capable of lifting the space­craft away from the booster in case of an emergency. “From the get-go, we tried desperately to put an abort system on the shuttle that would allow us to abort the crew and/or the orbiter off of a malfunctioning solid rocket or malfunctioning ssmes [Space Shuttle main engines],” Kraft said.

Originally we tried putting a solid rocket booster on the ass end of the orbiter, and the more we looked at that, the more we could not come up with a struc­tural aerodynamic qualification and weight that would accomplish that job. We looked at putting a capsule in the structure of the crew cabin, making it some­thing that would separate. We looked at the possibility of putting a capsule in the orbiter, at the structural problems of attaching a capsule, getting rid of the front end, making it strong enough, making it aerodynamically sound, building a control system that would allow it to descend under any and all Mach num­bers. And we decided if we do that, we can’t build a Space Shuttle. We cant af­ford the mass, and we don’t think we could build it in the first place.

So the answer to that question was, we will use the solid rockets that we have as our escape system and fly the orbiter back to the launch site if we have an abort. So we said to ourselves, the solid rockets have to, once you release them from the pad, bust those bolts on the pad, it has to be 100 percent reliable. And

Подпись: MAIN ENGINE Подпись: jORBITER Подпись: EXTERNAL jTANK

In the BeginningSPACE SHUTTLE VEHICLE

SOLID

ROCKET

[booster!

5. An early depiction of the Space Shuttle identifies major components as the orbiter, the three main
engines, the external tank, and the two solid rocket boosters. Courtesy nasa.

we always assumed it was, and any decision we made could not screw around the reliability of those solids. So our abort system was the solid rockets and a re­turn to launch site, rtls;you could fly that orbiter back.

Kraft said critics gave nasa a hard time about the shuttle not having an escape system. “I always thought that was unfair as hell,” Kraft said. “I don’t think they understood the system. And if you ask them over there [at nasa] today, I guarantee you won’t find five people who understand that’s what we did. But we did have an escape system, we had the solid rockets and the fly-back capability. Now, it didn’t save the Challenger, and nothing would have saved Columbia. But those two accidents were created by the fallacies of man, not by the machines.”

Recalled astronaut Charlie Bolden of the rtls abort:

While a lot of us flew a lot of them in [simulation], I’m not sure any of us ever believed that that’s something you really wanted to do. This was a maneuver in which something goes wrong shortly after liftoff, and you decide you’re going to turn the vehicle around and fly it back to the Kennedy Space Center. And the computer’s got to do that, so the software really has to work. It’s crazy, because

you’re going upside down outbound, and all of a sudden you decide you’re go­ing to go back to Kennedy. And while you’re still flying downrange, you take this vehicle and you pitch it back over so that it’s flying backwards through its own fire for several minutes. What has to happen is the computer has to calcu­late everything precisely, because it’s got to flip it over, have it pointing back to the Cape while it’s flying backwards, so that just before the solid rocket boost­ers burn out, it stops the backwards downrange travel and starts it flying back to the Cape. And then once that happens, then the solids cut off They separate; they go their way, and then you fly back for a few minutes, for another six min­utes, and the main engines cut off and you separate from the external tank. And that became a very tricky maneuver, because what you’re worried about was re­impacting with the tank, and if you did that, you were dead. So it’s a maneu­ver that. . . nobody ever wants to fly it, because just, it’s like, boy, this is really bad if you have to do this.

Once the general requirements were outlined from the mission base­lines and the general type of vehicle was decided, work began on figuring out how exactly to design a spacecraft that would meet the requirements. Making the process particularly interesting was the fact that the shuttle was a collection of very diverse elements that had to be designed to work as an integrated system. The orbiter, for example, ended up with engines that, by itself, it couldn’t use because they had fuel only when the orbiter was con­nected to the external tank. The diameter of the external tank is another example of the integrated approach used in designing the entire shuttle sys­tem, according to retired NASA engineer Myron “Mike” Pessin, who spent the bulk of his career working with the external tank. Taken as a single el­ement, there is no reason for the tank to have its 27.6-foot diameter. There was a constraint to the diameter of the solid rocket booster, however—it would have to be transported by rail from Utah to Florida, and so it was designed with a train’s dimensions in mind. That diameter determined the length of the boosters, which in turn established the range of locations for the explosive bolts that connected the boosters to the external tank. Given that engineers wanted to keep the connecting bolts off of the liquid hydro­gen and liquid oxygen tanks inside the external tank, they were able to es­tablish exactly how long the structure would need to be to make that pos­sible. Since they knew how much fuel the tank would need to hold, they could use the volume and the length to determine the needed diameter. Thus the diameter of the external tank was indirectly determined by the dimensions of the train that would be carrying the solids.

Another important part of the shuttle system design process involved computer technology that had evolved substantially since the development of nasa’s earlier manned space programs. “Now we get into the hard part of, okay, now we know the requirements, how do you make this all hap­pen?” Mattingly said.

And that all settled down certainly after Skylab, and maybe even after astp [the Apollo-Soyuz Test Project]. Then we started working. I remember Phil Shaffer was designated as the lead for pulling together all of our software and stuff. Be­cause the shuttle is such a highly integrated vehicle, it has the [software] archi­tecture that makes the system run, and then it’s got all of the applications which are the heart of the vehicle. And so we were building all of this from scratch, and in Apollo we were astounded we had computers. I guess Gemini had a lit­tle computer, and then Apollo had something which, by today’s standards, your wristwatch is far more powerful than what we had those days. But we were still astounded with what you could do with these things. Now we were going to build this shuttle with these computers and they’re going to be its lifeblood. There wont be a lot of direct wire. Everything goes on a data bus, and this was all relatively new for most of us.

It meant learning a whole new design process, and we learned that the software was the pacing item. We blamed it on software. When we think ofdeveloping soft­ware, we think of it as coding, “if/or " statements and counting bits, but in fact the massive amount of energy went center-wide into collecting the requirements— what does it have to do, write it down, and then see if you can package it, be­fore anybody could start worrying about building. That was an extraordinary operation. Phil drove that thing. I’m sure if Phil hadn’t been there, there would have been somebody that could have done it, but I have a hard time imagining anybody that could have done it the way he did. He just had the extraordinary personality and insight. He knew all the key players from the Apollo days, and they just set out and they went to work, and they really made the program go.

In spite ofall the delays that the shuttle program experienced—and we generally tended to blame that on truncated budgets, maybe some more money would have held the schedule a little better—the best I could tell, we were working as fast as that group of people [could]. It was such a massive job, and it just took so long to get everybody educated up to the same level, because it was all integrated. I don’t think when we started anybody knew that it was going to be such a challenge, and so we learned to do those things and went through it. This doesn’t sound like a CB [Astronaut Office] perspective, but. . . a little more than half[of the astronauts] were working the engineering side, working on the development ofthese things and trying to look ahead to see what was going to be required as part ofgetting started.

“We not only wanted to land on ten-thousand-foot runways, but we were going to be an airline,” Mattingly said, explaining that since the shut­tle would be a reusable aircraft with, ideally, a short turnaround time, nasa decided to turn to airline officials for help with how to do that.

So people went out and got contracts with American Airlines to teach us how to do maintenance and training, and we had people come in and start giving classes on how you give instructional courses and how we do logistics [in] the airlines. For a couple of years, we studiously tried to follow all that, and finally after a good bit it became clear that, you know, if there is anybody that’s going to ex­plain this to someone, it’s going to have to be us explaining it to ourselves. That’s where it evolved back into the way we had done things in the earlier programs.

Developing the systems was very much a group effort, Mattingly recalled.

I remember when we first started building the flight control schematics. Those are the most magnificent educational tools I’ve ever seen. I’ve never encountered them in any other organization. I don’t know why. I used to carry around a cou­ple of samples and give them to people and say, “This is what you really need." And they’d say, “Oh, that’s all very interesting," and then nothing ever seemed to happen. But working with people to put those drawings together, and then un­derstand what they meant and develop procedures and things from, was a mas­sive effort. During those days the Building 4 [at Johnson Space Center] and the building behind that, where flight control teams had some other offices, the walls were just papered with these things. People would go around, and they’d walk by it and look at it, and they’d say, “That’s not right. "They’d draw a little red thing on it and say, “See me. "And it was an evolutionary process going on continuously.

The shuttle was built with redundant systems. The idea was it should be able to suffer loss of any piece of equipment and still be able to fly safely. It was called “fail op, fail safe,” meaning that one failure wouldn’t affect nor­mal operations and that a second failure could affect the way the vehicle operated but not its safety.

That generally led to a concept offour parallel strings of everything. And that was great, but now how do you manage it, and what do you do with it? Now, a sche­matic has all of these four strings of things, sometimes they’re interconnected, and you could study those things, you’d pull those long sheets out, and you go absolute­ly bonkers—“Oh no. This line’s hooked to that. I forgot that." Trying to figure out how this all works. So you’d go get your colored pencils out, and you’d color-code them. By now the stack of these things is building up, and I’m really getting frus­trated in doing this dog-work job just before—I had to spend many, many hours for each drawing to get it sorted out before you were ready to use the drawing. So I said “We’ve got to take these things and get them printed in color, right offthe bat."

And so my friends in the training department said, “Well, you’re probably going to have to talk to Kranz about that. He’s not that enthusiastic about it." I thought, “Oh God." So I got an audience with Gene and went over and sat in his office and explained to him what we were doing in trying to get the train­ing program started and how we were trying to get ready to do that, and I re­ally wanted to get these things printed in color so that it would make it easier for people. I knew color printing would be a little more expensive, but it would sure save a lot of time. He said, “No. We’re not going to do that." I was just over­whelmed. I said, “Gene, why?" He didn’t say a word, he just turned and looked at his desk, and there on his desk, right in the corner, was this big mug filled with colored pencils. And he says, “That’s how you learn." And so that was the end of the story. I don’t know, I’ll bet today they’re still black and white. But that was Gene’s method of learning, and he figured that by having to trace it out, he had learned a lot, so he felt that others would benefit from that exercise. Even if they didn’t appreciate it, they would benefit.

The process of how the orbiter cockpit was designed would produce rath­er interesting and, in some cases, counterintuitive results, Mattingly said. He was part of a working group on controls and displays with fellow astro­naut Gordon Fullerton, which made decisions about the center console.

If you sit in the orbiter, the pilot and commander are sitting side by side in the center console. It was one of the few places when, if you put on a pressure suit, . . . you could see and touch. I mean, you can see the instrument panel. Stuffup here gets really above your head, gets really hard to see. It’s in close, so it’s diffi­cult for some of us older people to focus, and you cant see a lot. You have to do it by feel, which isn’t a good thing to do with important things. So the mobility was small, and this was prime real estate. We all knew it. As we went on with the program, every time someone said, “Oh, we’ll just put this here [in the cen­ter console]," we’d say, “No." We’d have a big office meeting. We’d all agree that, no, that’s not that important. We can put that here, we can do this. Well, after working on this thing for years, there’s practically nothing that’s important on the center console. We kept relegating everything to somewhere else, and it’s now the place where you set your coffee when you’re in the [simulator]. We protected that so hard, and poor old Gordo fought and fought for different things, and we’d think something was good, and then after we’d learn about what it really did and how it worked, we’d say, “No. You don’t need that."

Then there was the question of how the Space Shuttle would fly. Each airplane flies slightly differently, or feels slightly different to a pilot flying it, and the only way to really understand exactly how a plane flies is to fly it. Further, a pilot’s understanding of how airplanes fly is, to some extent, limited by the variety of airplanes he or she has flown. Those differences are rooted in the physical differences in the airplane’s control systems, a factor that means something entirely different with the computer-aided fly-by­wire controls of the shuttle. “There is a military spec that publishes about flying qualities, handling qualities of airplanes,” Mattingly said.

It started back in World War II, I guess, maybe even before. It tells you all of the characteristics that have to go into making a good airplane, like how many pounds of force do you put on a rudder pedal to push it. Well, even dumb pi­lots finally figured out that with an electric airplane this maybe isn’t really rel­evant. Then the engineers wanted to just throw out all of the experience and say, “Hey, we’ll just make it cool and you’ll like it." So we went on a crusade to rewrite this document, which turned out to be one of the most interesting proj­ects I’ve ever been in, because it required rethinking a lot of the things that we all took for gospel. Every airplane that a pilot flies is the Bible on how airplanes fly. Fortunately, in the office we had people who had flown a lot of different kinds of airplanes. But nevertheless, that shapes your image. And now you get into something that’s totally different, and there’s a tendency to want to make

this new airplane fly like the one you like the most. The software guys contrib­uted to this bad habit by saying, “Hey, it’s software. You tell us what you want, we’ll make it fly." I remember one time they gave us a proposal that had a lit­tle dial and you could make it a P-51 or a T-33 or a f-86 or a 747. “Just tell me what you want." We had a lot of naive ideas when we started.

While the computer for the Space Shuttle allowed many things that were groundbreaking at the time in the world of avionics, Mattingly pointed out that they were still quite primitive compared to modern standards.

I don’t remember the original size of the computer, but it had a memory that was miniscule by today’s standards, but it was huge compared to Apollo. By the time we finished this program, we had this horrendous debate about going to what we called double-density memory that would expand it. It was still nothing, and the only reason management did not want to change to it was for philosophic rea­sons. And IBM finally said, “Look, you guys said you wanted to buy off-the-shelf hardware. Let me tell you, you are the only people in the world with that version of a computer. So if you want to stay with the rest of the world, you’re going to have to take this one." And fortunately, we did, and still it was miniscule. Today I think they’ve upgraded it several more times so that it isn’t nearly the challenge. But that caused us to partition the functions in prelaunch and ascent and then get out of orbit and do some servicing things and then another load for reentry.

Don Peterson, who was selected as an astronaut in 1969 and flew one shuttle mission, said the orbiter computer systems were quite complicated.

My little desktop computer at home is about a hundred times faster and it has about a hundred times more capacity than the computers that were flying on the orbiter. They were afraid to change the computers very much because part of the flight control scheme is based on timing. If you change the computer, you change the timing, and you’d have to redo all the testing. There are thousands of hours of testing that have gone into there, and they know this thing works, and they’re very loathe to make those kinds of changes. They cant change the outside of the vehicle for the same reason; that affects the aerodynamics. So they can change some things in that vehicle, and they [improved] some of it. But they’re not going to make big, drastic changes to the control systems. It’s just too compli­cated and too costly. The flight control system on the orbiter is almost an experi­mental design. In other words, they built the system and then they tested it and tested it and tested it. They just kept changing little bits and pieces, primarily in the software, until it all worked. But if you went back and looked at it from a theoretical point of view, that’s not very pretty. You know what I mean? It’s like, gee, there doesn’t seem to be any consistent deep underlying theory here. It’s all patchwork and it’s all pieced together. And in a sense, that’s true. But that’s why they would be very loathe to try to make big changes to that, because put­ting all that stuff together took a long, long time.

Working on a project with so many systems that all had to be integrated but that were being developed simultaneously was an interesting challenge, recalled Mattingly. “Within the office, we were all trying to stay in touch with all these things going on in each of these areas to keep them some­what in sync from the cockpit perspective. So that gave us a lot of insight into all of these tasks that people were doing,” he said.

We even found, for instance, that as part of this development program, people working with thermal protections systems, the structure guys found that they were discovering limitations that were going to be imposed on the vehicle down­stream that we weren’t thinking about—if you fly in the wrong regimes, you will get yourself into thermal problems. Yet nothing in our flight control work or displays was considering that. We had never encountered anything like that before. So the guys, by working all these different shops, were picking up these little tidbits and we were trying to find ways to look ahead.

Another major change, Mattingly said, was developing and testing the flight control software for the shuttle. “We learned quickly that the man – machine interface is the most labor intensive part of building all this soft­ware,” he said, explaining that the code dedicated to computer control of the vehicle made up less of the software—and less of the time it took to develop it—than the code related to the interface that would allow the as­tronauts to control use of that software to control the vehicle. In addition, he said, a conflict arose because of the computer use needed to develop and test that software. To the engineers who were using those computers to de­sign the vehicle, the time the astronauts spent testing and practicing with the flight control software seemed like “video games.”

We ended up building a team of people: Joe Gamble, who was working the aero­dynamics; Jon Harpold, doing guidance; and Ernie Smith, who was the flight control guy. They all worked in E&D [Engineering and Development]. We all got to going around together in a little team, and we would all go to the simu­lators together, and we would all study things. We built a simulator from Apol­lo hardware that was called. . . its, the Interim Test Station. We had a cou­ple of people—Roger Burke and Al Ragsdale were two sim engineers that had worked on the cms [CommandModule Simulator] and the lms [Lunar Mod­ule Simulator]. They were very innovative, and they took these things before we had the Shuttle Mission Simulator that was back in the early part of the design and went to the junkyard and found airplane parts and built an instrument panel out of spare parts and had a regular chair that you sat in and had dif­ferent control devices that we had borrowed and stolen from places. These folks were so innovative; they could hook it all up.

“They took the initial aerodynamic data books and put them in a file so we could build something that would try to fly,” Mattingly said.

We even took the lunar landing scene television. In the Lunar Module Simula­tor they had a camera that was driven by the model of the motion and it would fly down over the lunar surface, and so you can see this thing, and that was por­trayed in the lms as what you’d train to. So they adapted that to a runway. We tried to build a little visual so we could have some clues to this thing, put in a little rinky-dink CRT [cathode-ray tube] so we could play with building displays. And we got no support from anybody. I mean, this wasn’t space stuff And it is probably one of those things I was most proud of, because we were able to get this thing into someplace where we could actually tinker with how were going to fly the vehicle and what we’re going to do and what the aerodynamics mean. It was only possible because we had these two simulator guys who were wizards at playing with software and this team from e&d who joined us.

We ended up realizing that we had built an electric airplane that had essen­tially only one operating flight control system. So we said, “Well, what if we’re wrong? No one has ever flown a Mach 20 airplane. This whole flight envelope is something that nobody’s ever had the opportunity to experience. So what do you suppose our tolerance is to this?" Because wind tunnel models for the as­cent vehicles, they fit in your hand, because the tunnels that were able to han­dle these things were small. The wind tunnel models for the orbiter were larger, but they’re still not all that big, and going through this tremendously wide flight regime where the air density is going from nothing to everything, and it’s just high speeds to low speeds, I said, “What’s the chance of getting all that right?" And yet as we played in these simulators, … we proved to ourselves that, boy, if you’re offon that estimate of the aerodynamics, you can often play with the soft­ware to make it right, but if the real aerodynamics and the software you have don’t match, it’s a real mess. I know I worried a lot about that.

So we came up with a concept that we would have some tolerances on the aerodynamics, and we would try to make sure that the flight control system could handle these kind of uncertainties in aerodynamics. We did something which is not typically done—we decided to optimize the flight control performance to be tolerant on uncertainties rather than the best flight control system they could build. The whole idea was, after we’ve flown and we have some experience and we know what the real world is, now we can come back and make it better, but the first job is to make ours as tolerant as possible to the things we don’t know.

While Mattingly was working with the computer models of the flight dynamics of the shuttle, astronaut Hank Hartsfield was on the other side of that research, working with the wind tunnel models and encountering the same concerns about the scalability of the data coming out of those tests.

As I recall, the shuttle program had over twenty-two thousand hours of wind tunnel time to try to figure out what it flies like. Because the decision had been made, there are no test flights. We were going to fly it manned the first flight, and an orbital flight at that, which demanded that, the best you can, [we] un­derstand this. Well, hypersonic aerodynamics is difficult to understand, the un­certainty on the aerodynamic parameters that you get out of the tunnel are big. The things that we were looking at in the simulations were if these uncertain­ties in the different aerodynamic parameters stack in a certain way, the vehicle could be unstable.

What we were looking for, for those combinations, statistically were possi­ble, but hopefully not very probable they’d happen, but if they did, that was the kind of things we had to plan for. It’s just an uncertain world. You can’t predict, because in the wind tunnel, you have to put in scaling factors. If you’re doing wind tunnel things off a small model, it doesn’t really scale to the big model per­fectly, and you have to make assumptions when you do that. The scaling ratios have a big factor, a big effect on what the real numbers are. So if you could fly a full-scale orbiter in the wind tunnel and it would go Mach 15 or something, it would be great, but you cant do that. You have a little-bitty model, and it’s a

In the Beginning

6. Space Shuttle vehicle testing in the fourteen-foot Transonic Wind Tunnel at nasas Ames Research Center. Courtesy nasa.

shock tunnel or something. You’d get a few seconds of runtime at the right Mach numbers and then try to capture the data off of that.

Astronaut Don Peterson was involved in studying the redundancy of systems on the orbiter, and particularly the flight control computers. In the report he pointed out that failure rates on some of the avionics could be high. On Apollo and earlier vehicles, nasa built “ultra-reliability com­ponents,” components that were overdesigned and tested to make failures less likely.

Failures on Apollo, for that reason, were pretty rare. But that’s very expensive. That’s a very difficult thing to do. I was told that after the lunar program ended, MIT had two of the lunar module computers left over, spares. So they just turned them on and programmed them to run cyclically through all the programs. I think they ran one of those computers for, like, fifteen years, and it never failed. It just kept running, and finally they turned it off They just said, “It’s not ever going to fail. ” That’s the way that equipment was built. But that makes it very

expensive. So when they built the Shuttle, they said, “We can’t do that. So what we’re going to do is, instead of ultrareliability components, we’re going to rely on something called redundancy. " They were going to have four computers, and they were going to have three tacans, and they were going to have four of this and two of that and so on. That way, you could tolerate failures. But as a result of that, the failure rate on some of that equipment was fairly high, compared to Apollo.

They also made the multiple units interdependent. “On a typical auto­mobile you have five tires, but that’s not five levels of redundancy because you need four of them,” Peterson explained.

So you can really only tolerate one failure. You can have one tire go bad and you can take care of that. But we got into that same situation on the shuttle because of the way they did the software. The shuttle, when it’s flying, the computers all compare answers with one another, and then they vote among themselves to see if anybody’s gone nuts. If a computer has gone bad, the other computers can over­ride its output so that it isn’t commanding anything. But to make that scheme work, you have to have at least three computers working. Otherwise, you cant vote. You could have [two systems voting], but if they vote against each other, you don’t know which one’s the bad one.

The decision was made to put five of the computers on the orbiter, with four of them active in the primary system, with the idea that this would create a system that could tolerate three failures. However, Peterson said, this produced much higher failure rates than expected. While the system provided a high amount of redundancy in theory, the reality was that be­cause of the way it was designed, the system actually could tolerate only one failure safely. The four primary computers were not truly redundant for each other; only the spare provided redundancy. If one computer failed, the spare would take its place. After that, however, further failures would endanger the cooperative “voting logic” between the computers that veri­fied the accuracy of their results.

But the complexity of the way the thing was put together kind of defeated the simplistic redundancy scheme that they had. It’d be like driving a car that had two engines or three engines, and any one of them would work. Well, that way you could fail two engines and you’d still drive right along. But if it takes two engines to power the vehicle, then you don’t have that, and if it takes three en­gines to power the vehicle, you don’t have any redundancy at all. It gets to be a game then as to how you trade all this off. When I looked at all that and we put the study together, we said, “You know, you’re going to have some failures that are going to really bother you because you’re going to lose components. ” For example, you’re on orbit and you’ve got four computers and one of them fails. Well, now you’ve got three computers left in the primary set. But do you stay on orbit? Because if you suffer one more failure, your voting algorithm no lon­ger works. Now you’re down then into coming home on a single computer and trusting it. And nobody wanted to do that.

So they said, “Gee, I’ve got four computers. I can only tolerate one failure, and then I’ve got to come home. ” We had four of some of the other components, and it was kind of the same sort of thing. If one of them fails, we are no lon­ger failure tolerant. We’ve lost the capability to compare results and vote, and so we don’t want to stay on orbit that way. So now, all of a sudden, the fact that you’ve got four of them causes more aborts because the more things you have, the more likely you are to have one fail. You’d get more failures and more aborts with four computers than if you’d gone with some other plan. That was pretty controversial for a while. We predicted—and there were some people that were really upset about that—we predicted a couple ofground aborts due to computer failures. Essentially we’d get chewed out for saying that, but in the first thirteen flights, we hit it right on the money. We had two ground aborts in thirteen flights.

When the shuttle was built, the air force was also using redundancy sys­tems, Peterson recalled. Then the air force built what it called confederat­ed systems, in which each component was independent. “They cooperated with each other, but they shipped data to each other, but they weren’t re­ally closely tied together,” Peterson explained.

The shuttle was tightly integrated. It runs on a very rigid timing scheme. The computers on the shuttle actually compare results about a little more than three hundred times a second. So it’s all tightly tied together. Well, when they decided to build the [International] Space Station, NASA said, “We’re not doing this in­tegrated stuff anymore. Boy, that was a real pain. We’re going to use a confeder­ated system. ” The air force, on their latest fighter, said, “This confederated stuff doesn’t work worth a damn. Were going to build a tightly integrated [system]. ” So they both went along for ten years or twelve years, and then they flip-flopped. The military’s going the way NASA originally went, and NASA’s now going the way the military went originally. I think the answer is, there is no magic answer to all that. Probably one concept is maybe not that much better than the other. It’s how you implement it and how much money you spend and how much to test. What do they say? The devils in the details. I think that’s right with all this stuff.

Mattingly recalled excellent cooperation between the engineering staff working on the shuttle and the Astronaut Office. “I seldom have seen that integration of the people that were going to fly it with the designers and people who were doing the theoretical work and the operators from the ground,” Mattingly said.

All of that stuff was converged in parallel, and I think that’s one of the reasons that the shuttle is such a magnificent flying machine. It does all the magic that we set out to do. I’m ignoring the cost because the shuttle, in my recollection, by the time it was sold to Congress, it was probably different than what the peo­ple in the trenches remember, but we had to do all these technical things, and it was a matter of faith that if you build it, it will be cheap. I mean, it was just simple. If you could reuse it, it saves money, and so you’ve got to make it reus­able. If you fly a lot, that will be good, and we’re going to fly this thing for $5.95, and we’re going to fly it once a week and that’s how we’re going to do this. And none of us were ever told to go build a vehicle that we could afford to own. And had we been told that, I doubt if we would have been able to do it. I think the job was so complex, you had to build one that flies in order to learn the lessons that say, "Now I know what’s important and what isn’t. ” I just think it would have been asking too much, but that’s just personal opinion, but it’s from hav­ing struggled through ten years of this development program. It was an extraor­dinary experience to do that.

The role of the Astronaut Office during the development of the Space Shuttle was quite different from what Mattingly experienced during the Apollo program. “Our involvement was far more extensive and pervasive, and a heck of a lot more fun,” he said.

I mean, this was really cool stuff. There was a problem every day, and you got to learn about all of these little things that were interesting. I spent a lot of time trying to understand the stress loads and the thermal characteristics on the tps [thermalprotection system], and how do you get it to stay on, and all of those things were things that came through the office as experiences that really were just extraordinary opportunities to go see that. As we moved down the stream and we got into some of these development programs and started turning out hardware, we started splitting people up to go follow different components of hardware, whether it be the engines or the SRBs or the orbiter.

The decision to have the orbiter be an unpowered glider rather than a jet during its return to Earth and the various ramifications of that decision were also among the things that had to be considered during development. “Some­where earlier in this development stage, we went through a series of activities where the first orbiter was going to have air-breathing engines, and it had some solid rockets that were on the back that were for aborts,” Mattingly said.

Right off the pad you could fire these two big rockets, and they would take you off in a big loop so you could come back and land. We had these air-breathing engines that were going to—after you come down through the atmosphere, you open the door and these engines come out, and you light them and you come around and land. They had enough gas for one go-around. The other thing we had was the big solids were to have thrust terminations and ports that blew out at the front end so you could terminate thrust on them if you needed to in an emergency. Every one of those devices was something which had a higher prob­ability of killing you by its presence than it would ever have in saving you. I’ll put that ejection seat in the same boat. Everybody was willing to get rid of the air-breathing engines. They were really, really not a very bright idea. And we got rid of the thrust termination and we got rid of the abort solid rockets. My guess is John Young was probably the most active stimulus in pushing those is­sues, and that was one of those cases where the flight crew perspective and the engineering perspectives converged. We all wanted to get rid of these things, and yet we retained the ejection seats for reasons which I will never understand. If anyone knew what the useful envelope of those ejection seats was and the price we paid to have them. . . . But it had become a cause: “You will protect these kids by giving them an ejection seat." So we had one, not that anybody wanted to ever use it, but it was there.

Astronaut Bonnie Dunbar was still an undergraduate student at the Uni­versity of Washington during the early portions of shuttle’s development, and she worked with the school’s dean of ceramics engineering, who had received a grant to work on the tiles for the shuttle’s thermal protection sys-

In the Beginning

7. A worker removes a tile as part of routine maintenance activities on the orbiter fleet.

Courtesy nasa.

tem. nasa’s earlier manned spacecraft had used ablative heat shields, which absorbed heat by burning up, protecting the rest of the vehicle. Such a sys­tem was simple and effective, but for the new, reusable Space Shuttle, nasa wanted a reusable heat shield, one that could protect the vehicle without itself being destroyed. The solution that was settled upon involved a vast collection of tiles and “blankets” covering the underside of the orbiter and other areas of the vehicle that would be exposed to extreme temperatures.

“First of all, tiles are a ceramic material, so by definition they’re brittle,” Dunbar said.

But the reason they have an advantage over metals is that they don’t expand ten times over their thermal exposure range. It’s called the coefficient of thermal ex­pansion. Also, they are an insulator; they don’t conduct heat. We looked at met­als, or what they call refractory metal skins, and there are two disadvantages. You still have to insulate behind them, because metals conduct heat. The other is that when you go from room temperature, let’s say seventy-five degrees Fahr­enheit, to twenty-three hundred [degrees], you have a large growth. It’s like your cookie pans, I guess, in the oven. So the airframe would distort. The ceramic materials [have] very small thermal coefficients of expansion, ten to the negative sixth, so you’re not going to see a lot of deformation. Also you could, on a very

low density tile, expose the surface to twenty-three hundred degrees Fahrenheit, and the backface, three inches deep, would not see even close to that, less than a couple hundred degrees, till after you’re on the ground. It’s a very slow coeffi­cient of thermal expansion and heat transfer. So ceramics had a definite advan­tage. We knew that from the work we’d done in the sixties, and in fact, ceramics were already being used as the heat shields on nose cones for missiles and so forth. So the next big challenge was to put them in a low-density, lightweight form that could be applied to the outside of a vehicle. Apollo vehicles, Gemini, Mer­cury, were all covered by ablators, which meant that they burned up on the re­entry to the Earth’s atmosphere and could not be reused. The tiles were meant to be reusable. They didn’t deform. They didn’t change their chemistry. We had to, though, shape them so that they were the shape of an airplane, so we had all the aerodynamic features there. So we sort of did a little reverse engineering, in that we said, “Okay, here’s what the shuttle looks like; got to maintain that shape. Here’s how hot it gets from the nose to the tail. Most of the heat’s at the nose, on the nose cone, and the leading edges of the wings. We want to make sure the aluminum substructure doesn’t get over 350 degrees Fahrenheit; that’s when it starts to change shape. So how thick does the tile have to be?" So we used all those limits and constraints, then you’d use the computer. . . to calculate how thick each tile had to be. Then we started looking at, well, okay, how big should each tile have to be? Could I just put large sheets of tile on there?

Well, we started looking at what the structure does during launch, and now we’re getting to something called vibroacoustics. There’s a lot of force pressure on the vehicle, a lot of noise, if you will, generated into the structure, and it vi­brates. We calculated that if we put a foot-by-foot piece of tile on there, the vi­bration would actually break it up into six-by-six-inch pieces. We said, “Well, we’ll design it six by six. " So you’ll see most tiles are six by six. Now how close do you put them? We thought, well, you cant get them too close, because dur­ing that vibration they’ll beat each other to death, because they’re covered with a glaze. You’ve got silicon dioxide fibers that are made into very low mass tiles, nine pounds per cubic feet, or twenty-two pounds, and to ensure they don’t erode in the airstream when you reenter, they’re covered with a ceramic glaze. So that’s also brittle, so you can’t get them too close or they’ll break the glaze. You can’t get them too far apart or, during reentry, the plasma flow will penetrate down in those gaps and could melt the aluminum. So that’s called gap or plasma in­trusion. So that then constrained what we called the gap. Then from tile to tile, how high one was compared to the next one, we called step. That became im­portant because if you had too large a step towards the leading edge of the wing, that would disturb the boundary layer, and you would go up the plasma, and instead of having smooth layers, it would start to transition to turbulent, from laminar to turbulent, and turbulent results in higher heating. So that controlled the step. So gap and step were very important to that as well.

“Those were all challenges,” Dunbar said. “We depended on advances in computerized machining capabilities, wind tunnel work with models to help us determine the requirements on step, the manufacturing, just every­thing. Firing a tile, a certain temperature and time was important to main­taining its geometry. . . . It’s, I think, a real tribute to the program that if you look at follow-on programs, even in NASA but also in Japan or in Eu­rope or even the Russians, who built the Buran [Soviet shuttle], you’ll find that the system on the surface is very similar to the shuttle tile system. It was a good solution.”

Dunbar said that working on the shuttle during that early time was an exciting opportunity.

This was the next-generation vehicle. Not only was it next generation, it was…

“transformational" is the word we use now. If you think about it, everything to that point was one use only. Couldn’t bring any mass back. We sent a lot of things into orbit that we had to test and leave there, and it became a shooting star, coming back to Earth. So this transformed our ability to do research. It’s why we have a space station now. We not only learned from Skylab, but we flew [on] Spacelab countless research projects that we could bring back to Earth, get the results out, diagnose problems with equipment. I think it saved the govern­ment billions of dollars, because we didn’t throw it away each time. So it was exciting, and we knew what it could do. New technology. It was leading edge on not only the thermal protection systems, but it was the first fully fly-by-wire vehicle, in terms of the computers and the flight control system. The main en­gines were also a pathfinder as well, and so it was exciting, even if it delayed till ’81. If you think about it, we baselined it to the contractor, to Rockwell, in 1972, I believe. So nine years later we have a vehicle, a reusable vehicle, flying.

Astronaut Terry Hart was the Astronaut Office’s representative in the de­velopment of the Space Shuttle main engines.

Since I had a technical background, mostly mechanical engineering, John Young had asked me to follow the main engine development. This was a couple of years before sts-1. In fact, it was ironic that we showed up [as NASA astronauts] in ’78, and everyone said we’re one year away from the first shuttle launch, and two years later, we were still one year away from the first shuttle launch, and it was really because of two main areas of technical difficulty. The main engine development was somewhat problematic, with some turbo pump failures that they’d had on the test stand, and the tiles. We had difficulty with the tiles be­ing bonded on properly and staying on. But the main engine was one that John Young wanted me to follow for him, and so I spent a lot of time going back and forth to [Marshall Space Flight Center in] Huntsville [Alabama] and to nstl, the National Space Technology Laboratories, in Bay St. Louis [Mississippi, cur­rently called the NASA John C. Stennis Space Center], where NASA tested the en­gines. And Huntsville, of course, was where the program office was for the main engines. And that was very exciting. I mean, I was like a kid in a candy store, in the sense that a mechanical engineer being able to kibitz in this technology, with the tremendous power of the fuel pumps and the oxidizer pumps, and the whole engine design, I thought, was just phenomenal. The hard part of that job was when we had failures on the test stand, which were, unfortunately, too fre­quent. I’d get the pleasure of standing up in front of John Young and the rest of the astronauts on Monday morning to explain what happened. And, of course, everyone was always very disappointed, because we knew this was setting back the first launch and it was a jeopardy to the whole program. But we got through that, and the engines have done extremely well all through the program here, where it was always thought to be the weak link in the design.

Astronaut Don Lind was involved in the early planning and development of the remote manipulator system, the shuttle’s robot arm.

I guess the first significant assignment I had [for the shuttle] was in develop­ing the control system for the remote manipulator system, the RMS. In the hinge line of the cargo bay doors, there is an arm that’s articulated pretty much like the human arm. It’s about as long as two telephone poles, and it’s designed for deploying and retrieving satellites. Again, somebody had to worry about the op­erational considerations of that arm. It was built by the Canadians with the agreement through the [U. S.] State Department, and I was assigned to work on that. So I made a lot of trips into Canada to work with those people. The peo­ple who were actually building the hardware were very, very compatible, very easy to work with, and we had a very nice working relationship.

Lind contributed to the development of the three different coordinate systems that were going to be built into the arm’s software.

One coordinate system, obviously, applied when you’re looking out of the win­dow into the cargo bay, and so you want to work in that coordinate system. If you wanted the arm to move away from you, you pushed the hand controller away from you. Also, if you’re trying to grasp a satellite up over your head and you’re looking with the TV camera down the fingers at the end of the arm, which is called the end effector, and you want to move straight along the direction the fingers are pointing, you don’t want to have to try to figure out which way you should go, so you shift to a totally different coordinate system. So if you’re look­ing in the TV picture with the camera that’s mounted right above the end effec­tor, you want to push the hand controller straightforward. You want it to move straight forward in the television picture.

Lind also helped answer the question of how the hand controllers were to be configured.

We wanted hand controllers where the translation [movement] motion would be done by one hand controller, which we decided would be the left hand, and the rotational motion controlled by a hand controller which would be handled with the right hand. We decided, as a joint decision, that the hand controller for translation should be a square knob.

Then I said, “Now, remember you’re floating. You’re floating, so you’ve got to hang on to something while you’re translating, and you don’t want your bobbing around to affect the hand controller. So you need to put a square bracket around it so you can hold on to the bracket with your little finger and can use the hand controller." “Oh yeah, we hadn’t thought of that. Well, how big do you want it to be?" We actually measured my hand and designed the controller and bracket to the physical dimensions of my hand. Obviously, when you make a decision like that, then you have five other astronauts check it out, and they say, “Yeah, that was a really good decision." I didn’t want the hand controller for the right hand to be mounted square on the bulkhead, because the relaxed position of your arm is not at a square angle; it’s drooping down to the side. And I wanted that position to be the no-rotation position. We set up a simulation, and I stood up there, and they measured the angle of my arm and then built a bracket to mount that hand controller just exactly the way my arm relaxed. And again, we had several other astronauts check it, and they said yes, that was a fine thing. So the hand controllers were literally fine-tuned to my design.

Other people were worrying about the software, how to implement these co­ordinate systems. Other people were doing all the very sophisticated engineering. But the human factor was my responsibility, and basically it was a very pleas­ant experience to work with the Canadians, with one exception. The arm has two joints: like the elbow, and like the shoulder; one degree offreedom in the el­bow, two in the shoulder, and three degrees offreedom in the wrist, so there are three literal components to the wrist junction. They had mounted the camera on the middle one. As you maneuver in certain ways, the wrist has to compen­sate for the rotations of the other joints, and every once in a while the TV pic­ture would simply rotate. Not that anything had actually rotated, but the wrist was compensating. I said, “That’s unacceptable. ” They said, “No, no, no, no, it has to be there. That’s the cheapest place to put it. ” The engineers were all in agreement that this was a mistake, because you could lose a satellite when sud­denly the picture rotates and nothing really has happened. But the management people said, “This meets our letter of intent with the State Department. We’re not going to change it. ” So in one meeting I had to be very unpleasant. I said, “Now, gentlemen, if we ever lose a satellite because of this unnatural rotation, I will personally hold a press conference and say that you had been warned, and it’s the Canadians’ fault. ” They looked at me like, “Ooh, you’re nasty. ” At the next meeting, they said, “Well, we’ll change it, and it doesn’t cost as much as we thought in the first place.” Usually you could get good cooperation, but occa­sionally, particularly with people up in the bureaucratic levels, you had to be a little bit pushy. I try not to be pushy, but that’s one time I did.

Astronaut George “Pinky” Nelson was involved in the development of the Extravehicular Mobility Unit (emu), the spacesuit used for conduct­ing activities outside of the spacecraft. “The suit was one of the long poles in getting the shuttle ready to fly,” he said.

The folks in Houston who were in charge of it, [Walter] Guy and his group, were really working hard, and it was a difficult task to get it pulled together. The suit actually blew up shortly before sts-1. I was home working in my gar­den. I was playing hooky one afternoon, and I got a call from George Abbey.

He said, “Where the hell are you?” “Well, I’m home working in the garden.” He said, “Okay. Get in here. We just had an accident with the spacesuit. "They were doing some testing in one of the vacuum chambers in Building 7, and they had the suit unmanned, pressurized, in the vacuum chamber. They were going to do some tests and they were going through the procedures of donning the suit and flipping all the switches in the right order and going through the checklist. There’s a point when you get in the suit that you move a valve. There’s a slider valve on the front of the suit, and you move this slider valve over, and what it does is it pushes a lever inside a regulator and opens up a line that brings the high-pressure emergency [oxygen] tanks on line. You do that just before you go outside. You don’t need them when you’re in the cabin, because you can always repressurize the airlock. When you’re going to go outside, you need these high – pressure tanks. They’re two little stainless steel tanks about six inches in diam­eter, maybe seven. And it turned out that when this tech did that, he threw that switch and the suit basically blew up. I mean not just pneumatically, but burst into flames [and] got severely burned. It was pure oxygen in there. The backpack is made basically out of a big block of aluminum, and aluminum is flammable in pure oxygen. So this thing just went “whooff,” went up in smoke.

So then I was put on the Investigation Board for that, and spent I don’t know how long, a couple months at least, just focusing on what had caused this and could we identify it and fix it and get it ready so that it wasn’t the long pole for flying sts-1. So I learned even more about the design and manufacturing and materials and all ofthat in the suit during that process. It was fascinating. And the NASA sys­tem for handling that kind ofan incident really is very good. We’ve seen it with the big accidents we’ve had. They really can get to the bottom of a problem very well.

After that, Nelson said, there weren’t any major problems in the devel­opment of the suit. “There were lots of little stuff. The displays and con­trols on the suit are a challenge because, one, you have to see them from inside the suit, looking down, so a lot of these old guys in the office who were, you know, the stage I am in my life now, where I have to wear read­ing glasses, couldn’t read the displays because they were close to your face. So we worked on lenses and all kinds of ways to make the displays legible to people with old eyes.”

For all the capabilities built into the vehicle, one of the notorious dis­appointments of the Space Shuttle program is that launch costs ended up being much higher than promised. The original appeal of the shuttle was that its reusability would bring launch costs down dramatically, but those dreams were never fully realized. Explained Don Peterson,

The shuttles, unfortunately, are pretty difficult to work on. When the military builds an airplane, it tries to make everything in the airplane designed so that you can remove and replace parts quickly and easily. The shuttle is much more difficult to get to some of the stuff. Therere not big [easily opened]panels on it. You cant release a few latches and open a big panel on the side of the orbiter. You literally have to take it apart to get into it. You can go in through the in­side, through the bay, and get to some of that stuff, but even then you’re removing parts that aren’t designed [for that]. It’s not like opening doors and looking inside. The military builds a lot oftheir stuff to be easy to work on, and they really didn’t build the shuttle that way. So the shuttle is more expensive to operate. For exam­ple, the little jet engines, there’s, like, thirty-eight of them, I think, on the orbiter that control attitude when it’s on orbit. If one of those engines fails, you cant just unscrew some things and take it out. You have to cut it out with a torch, and you have to weld the new one back in, because they didn’t build it to be removed. The heat shield is [24,300] little individual tiles, and they’re all different shapes and different thicknesses, and so every tile is like a little individual item. When the shuttle comes back, they have to inspect visually, and with a pull device, every single tile. If any of them don’t pass, you’ve got to cut that one out and clean off the glue and go get the new one and put it all back. Those are very high mainte­nance items. So the shuttle really wasn’t built to be easy to maintain, and that’s because NASA has always had, as [former Johnson Space Center director] Gerry Griffin used to say, a standing army at the Cape that did all that, and nobody really worried about it. If you needed something done, you just called and they sent over four or five guys and they fixed it. But that’s expensive.

The shuttle was designed to fly, I think it was fifty flights a year, and they were going to have five shuttles to do that. So each shuttle would fly ten times in a year. Well, right now the whole fleet’s only flying about eight times a year. Well, you’re trying to amortize the cost of the whole program over eight flights. It’s like we’ve got all this capability to repair and replace and analyze and monitor things, and we’re not using a whole lot of it. If you were flying fifty times a year, the cost per flight would go way down because you wouldn’t add that much to the facilities and the maintenance costs. The facilities costs don’t change much if you never flew. You’ve still got to have all the facilities, and you’ve got to pay for all that. You have to keep this whole group of special­ists on, technicians and people, to do the work. With eight flights a year, some of those guys may only get used twice a year, but you’ve got to pay them and you’ve still got to have them there. If you were flying a lot more, the cost per flight would go way down.

George Mueller, the NASA head of human spaceflight who launched the Space Shuttle program, explained that there were several factors that drove the operational cost of the shuttle up, including many decisions, like the use of solid rocket boosters, that reduced development costs at the outset and pre­sented Congress a lower buy-in budget request to build the vehicle but that resulted in higher operational costs once the shuttle started flying. Howev­er, he said, the ultimate problem with the shuttle was that it ended up being designed to use far more people to process it than were absolutely necessary. “If you really want to know why the shuttle failed, it’s because they designed it to use all the people from Saturn and Apollo, to keep them employed.”

Countless technical problems had to be overcome, and ultimately the shuttle’s greatest limitation was that it was designed to be too nice.

Former jsc director Chris Kraft, however, still speaks highly of the shut­tle. “It’s the safest spacecraft we ever built.” Kraft noted that while shuttle crews have been lost because of problems stemming with the solid rocket boosters and the external tank, the orbiter itself has not been responsible for any fatal accidents. “The orbiter itself is flawless, since we’ve been flying. Absolutely flawless.” Rather than retiring the shuttle, Kraft argued, NASA should have continued to make it better and continued to fly it, adding that many ideas for improving the orbiter were never implemented. “That’s what we should still be doing. We still ought to be improving. We could improve the hell out of it. We could improve the hell out of the thermal protection system, we could improve the control systems, get rid of the apus [auxilia­ry power units]. All of that has been designed and is ready to be built. You don’t have to stop and redesign it, it’s done.”