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.