Eyes Turned Skywards

Part IV, Post 17: Orion Expedition 1
  • Good evening, everyone! This week, I'm pleased to bring you a special treat. Today, we're finally following the crew of Orion Expedition 1 on their way to their mission in the first semi-permanent outpost on the Moon. However, Workable Goblin and I wanted to make sure you all could get a feel for life onboard the outpost and make it more than "just another mission," so we turned to someone who's a past contributor to the TL, and better with narrative writing than seems quite fair given his other talents. Today's post is thus brought to you by none other than nixonshead, and I hope you'll all enjoy this as much as I did!

    Eyes Turned Skyward, Part IV: Post #17

    This is it!

    Edward Boxall, ESA astronaut, moved his booted left foot down to the final rung of the ladder and leaned backwards. Above him, the bulk of Clarke’s Descent Stage loomed brightly, contrasted against the black sky. No starlight made it through the layer of golf filter on his helmet visor, and the Earth, huddling close to the polar horizon, was hidden out of sight behind the lander.

    Leaning to look downwards, Ed could see dark streaks in the grey gray lunar regolith where the lander’s rockets had exposed the underlying bedrock. Small stones scattered around the landing site cast long shadows in the almost horizontally slanting sunlight, whilst the bootprints of his crewmates looked like black holes in the world. Soon, his own bootprints would be joining them.

    “Hey, Ed, you gonna join us today?”

    Ed smiled at the voice of the Mission Science Officer, always adept at breaking the tension in any situation.

    “On my way, Winch,” Ed called back over the radio. “Just one small step…”

    Ed’s right foot pressed into the gritty surface, quickly followed by his left.

    “For the people of the United Kingdom, the member nations of the European Space Agency and the world, may our mission here herald the beginning of a new phase of human exploration and habitation on our nearest neighbor.”

    Mission Day 6: Home Sweet Home

    After the excitement of yesterday's landing and our trek over to the Orion hab, today we awoke for our first morning in our new Lunar home. Unlike on earlier Artemis missions, this hab will go on to host other teams on future missions, so we need to take special care to keep the place clean and tidy. Right now it still has that 'new car' smell.

    Compared to my two previous missions on Freedom, I've already noticed several differences as well as a few similarities. By far the biggest difference is, of course, the gravity. A big benefit of this is when you put something down, it stays put. On Freedom I was forever losing pens, eating utensils, tools and other nik-naks that would drift into obscure corners the moment you let go of them. On the other hand, once I'd gotten used to it, my zero-gee sleeping bag was more comfortable to sleep in than the lightweight bunk in my miniature cabin in the hab's inflatable "loft". The low gravity means it's all too easy to bounce out of bed when rolling over. Hopefully, I'll soon be proficient enough to get a full night's sleep without risking a fall.

    Another bonus of having gravity is the ability to use proper cups for drinking from rather than the squeeze bottles used in microgee. Again though, you have to be careful not to slosh your beverage over the edge of the cup in a low gravity swell. Anne lost half her celebratory Tang last night toasting our arrival a little too vigorously.

    This morning, Winch and Anne tried out another new innovation for the Orion programme when they made the first EVA in the new hard-shell Moon Suits. These are officially named the "Articulated Lunar Excursion Suit", abbreviated to sound like "Alice" (at least to my British ears), but we usually refer to them as "Turtle Suits". Unlike the older A9L model suits we used to come across from Clarke yesterday, we can enter and exit the suits through a hatch in the backpack without needing to use the main airlock. This reduces the loss of breathing gasses associated with cycling the airlock, as well as cutting down on the amount of dust we track back into our living area.

    Winch and Anne stayed out for just over four hours in the suits, and tell me that they're much easier to work in than the old suits, although Winch had a few problems lining up for re-docking. There are handrails to help guide you in as you back up to the dock, but Winch still needed three tries to click into place. Anne got it first time - pilot's reflexes, she said. According to the schedule from Houston, Phil and I should get our chance to try out the suits tomorrow, deploying experiments around the base site. The real fun though will come later in the week, when we unpack the pressurised rover from the cargo lander. I can hardly wait for our first test drive in the new wheels!

    "One-two, one-two. How're you reading me, Anne?" Ed Boxall, the first British astronaut on the Moon, tapped the side of his Snoopy cap experimentally as he stood in his Thermal Control Garment next to ALES-1's suit lock.

    "Loud and clear, Ed," Anne Holcomb confirmed from the other side of the room. "Are you getting me through your headphones?"

    "Affirmative, I've got good reception," Ed reported. "How about you, Phil?"

    Mission Commander Phil Whitt, similarly attired to Ed, gave a thumbs up as he replied. "Also a good signal. Let's hope it's just as good at a range greater than five feet!"

    "Anne and I had no problems yesterday," Mission Science Officer Winchell Chung told the pair. "It got a little scratchy when relaying through Mesyat, but line-of-sight was clear as a bell. In any case, you're not going far today, so if you run into any problems just wave at a camera and we can come get you."

    On the communications workstation, Holcomb double-checked her read-out before reporting in to Mission Control. "Houston, Orion. Please confirm you have good Alice-1 and -2 comms relay, over." Following a brief but noticeable lightspeed delay, the voice of Capcom came back through both the console speakers and Boxall's and Whitt's headsets: "Orion, Houston, that's a roger on our side, we have good signal on both Phil and Ed. You have a go for EVA at your leisure."

    "Okay guys, time to get your shells on," said Chung. "Just like in training, you trigger the hatches and I'll confirm the seal before you undock, okay?"

    "No problem," said Whitt as he grabbed the handhold above the suitlock hatch and swung his legs into the waiting suit. Ed followed suit, easily managing to lift himself into position in the 1/6th gravity of the Moon. Like Phil, Ed slid in legs first through the open backpack of the Articulated Lunar Excursion Suit, slipping into the suit's legs before pulling in his arms and pushing them through the holes in front of him until his fingers hit the ends of the gloves. "This reminds me of putting my son into his romper suit when he was a toddler," Ed commented as he pulled his torso and head fully into the suit.

    "You don't squirm around half as much as my daughter did when I tried to dress her!" Chung replied. "Okay, Phil, you look good in there. Go ahead and close the backpack."

    "Okay Winch," Phil responded. Ed heard a few clicks through the open back of his suit, but otherwise just hung in place waiting for Phil to get to him. With the protective cover still over ALES-2's helmet, there wasn't even a view to enjoy yet. As he waited, Ed started to hum randomly.

    "Jeez, Ed, you didn't take that blasted kazoo into the suit did you?" came Holcomb's plaintive question.

    Ed laughed. "No, that's pure Edward Boxall, unplugged!"

    "You'd better not have anything unplugged," Chung put in. "Otherwise this EVA is liable to be scrubbed". The American astronaut was now behind Ed in the cabin. "Nope, looks like you're all wired up as you should be. Ready for seal?"

    "Ready," replied Ed. He triggered the closure mechanism, and with a quiet "thunk!" the background noises of the Orion cabin, which Ed hadn't even noticed up to now, abruptly ceased. Alone in the dark, with just the low whirl of the helmet fan, Ed was unpleasantly reminded of the sensory deprivation tests he'd undergone at Cologne when he'd first been selected as an astronaut. This time though, the silence didn't last long as Holcomb's voice came through his headphones: "Alice-1, Alice-2, Orion. Comms check."

    "Alice-1 comms okay," came Whitt's voice, before Ed responded "Alice-2, comms are good."

    "Okay guys," Holcomb replied, "It's all looking good from here. Lift up your dust covers and stand by for undock.

    "Roger, lifting cover now" Ed reported as he slowly moved the stiff, unresponsive arms of his moonsuit to bring his gloved right hand up to the fabric covering over his helmet. Moving carefully in the unfamiliar suit, he pushed wire hoop attached to the cover upwards and looked out upon the harsh, raw beauty of the lunar surface. Just as it had on his earlier EVAs in the old pressure suits, the view took his breath away, and he sent up a silent prayer of thanks to God that he was lucky enough to be alive at a time when such miracles were possible that he could walk upon the Moon.

    "Visual on Alice-2," came Whitt's voice through the radio, and Ed turned his head to see Phil's suit clinging by its backpack to the cabin wall next to him. Whitt's sun visor was down, so Ed couldn't make out his face, but the other astronaut gave him a cheery wave of recognition. Ed returned the wave. "Hi Phil, fancy meeting you here. It's a small world, eh?"

    "Smaller than Earth, that's for sure!"

    Just then Holcomb's voice broke in over the radio. "If you two have finished your comedy routine, we’ve confirmed a good seal on our side. You can undock when you’re ready."

    "Thanks, Anne," Whitt replied. "Alice-1 undocking... now!"

    Ed watched as Whitt's suit jerked forward, the rear half of his backpack emerging from the recessed suitlock as he held on to the twin railings either side of him for balance. "I have a good separation," Whitt told Holcomb. "I'm free standing on the platform."

    "Roger, Phil. Ed, ready to go for Alice-2 undock."

    “Roger.” Ed forced the stiff gloves of his suit to grip the side rails and pulled sharply forward. There was a brief resistance and a loud click as the backpack disconnected from its berth. “I’m out,” Ed called through the radio. Wow, I really am, he reminded himself. Outside on the Moon...

    Standing at the edge of the platform that topped Orion’s descent module, Ed looked out across the bleak landscape in awe. If he leaned over the railing and looked downwards (something that was much easier in his articulated Turtle-suit than it would have been in the old A9L), he could see the scars where the base’s descent had disturbed the top layer surface dust. A tangle of footprints and wheel tracks surrounded the lander, along with various boxes and equipment unpacked by Holcomb and Chung on their previous EVA, but out beyond a couple of hundred meters the ground was undisturbed. Primordial. Where no man has gone before...

    “Hey, Anne,” Whitt’s voice came over the radio. “Kill the lights for a second, will ya?”.

    “What’s that?” Ed asked nervously, tightening his grip on the handrail.

    “I just want to try something out,” came the commander’s enigmatic reply.

    “Okay, Phil,” Anne called from inside. “Shutting down the floods now.”

    As the hab’s external lights winked out, Ed found himself plunged into darkness. Both the sun and the Earth, neither of which ever rose more than a few degrees above the horizon here, were on the other side of the hab, so the only natural light was that reflected from the dusty surface of Shackleton Crater’s rim. Ed moved to switch on his helmet lamp, but Whitt said “Hold on a second. Push up your sun visor and let your eyes adapt.”

    Ed did as he was told, sliding the gold-coated visor upwards. Looking over to the silhouette of Whitt in Alice-1, Ed could see the commander using his arms to shield his eyes from the surface moonlight, his body tilted backwards.

    “Just take a look at that,” Whitt breathed.

    As Ed’s eyes turned skyward, he finally saw why Whitt had ordered the lights off. Shielded from the glare of the sun, with the bulk of the lunar surface hidden from view by the Hab’s descent stage and unfiltered by his protective visor, the full glory of the night’s sky could at last be seen.

    “My God. It’s full of stars…”

    Mission Day 10: Faster. Higher. Stronger.

    Just like many of you back on Earth, we four here at Orion are looking forward to this summer’s Olympic Games, and yesterday Anne and Phil took part in the very first Lunar Games! Unfortunately their heavy Turtle Suits meant that the Long Jump and High Jump events didn’t set any new records, despite the gravity being 1/6th of what you have on Earth. The Weightlifting event also fell a little short of championship contention, despite Anne carrying an impressive 25 kg of samples for sorting and analysis in the course of her moonwalk (though not all at once).

    The highlight of course was Phil Whitt’s participation in the first Earth-Moon Olympic Torch Relay. The torch, specially designed to withstand the rigours of space travel, is the same one that visited Freedom back in March, so it’s already a seasoned space traveller. We brought it up with us in Clarke, packed in a protective container that we carried over to Orion on our first day. Yesterday, Phil opened the box and carried the torch on a lap of the Hab, with some gold insulation foil doing duty for the flame. (A real flame of course is not possible on the airless Moon, and even if it were mission rules would forbid carrying the torch’s flammable fuel with us).

    After its lap of honour, Phil packed the torch back into its box, which we’ll carry with us when we return home. It will then re-join the terrestrial relay race (this time proudly alight) and be used at the opening ceremony on 8th August. A fitting symbol of international cooperation on Earth and in space!

    After two days in the pressurized rover’s tiny cabin, Ed Boxall was longing for the wide, open spaces of the hab module. He and Anne Holcomb had driven almost sixty kilometers around the rim of Shackleton crater, mostly sticking to the peaks but occasionally dipping into the shadowed depression of the crater itself. Ed had never enjoyed road trips, and especially disliked camping. But with no motorway service stations or roadside motels within a quarter-million miles, the camped two-person cabin had to do duty as cockpit, bedroom and bathroom all in one. The dust - which had gotten into the cabin somehow despite the use of suitlocks - was bad enough, but the smell… Maybe I should try describing the smell in my next blog entry, Ed thought to himself. Just to see if the PR guys in Houston let it through…

    As if the thought was enough to summon them, the radio crackled into life. “Rover, Houston, how do you read?”

    Holcomb toggled the set and replied “Houston, rover, we read you fine. We’re about four kilometers out from Orion and heading back.”

    “That’s our estimate too, Anne,” the response came a couple of seconds later. “Ah, we have a request from the science team to make another small diversion.” Even from this distance, Ed could hear the nervousness in the CapCom’s voice at raising the topic. “They’ve identified another potential LITT site close to your track and, ah, they’d appreciate a little ground truth survey.”

    Holcomb rolled her eyes and turned to Ed. “Another survey! You want this one?”

    “Hey, I got the last one!” he protested.

    “Yeah, but I got the two before that, plus I’m designated driver for the next two hours.”

    “Rover, Houston, did you read my last?”

    “We heard you, Houston,” Anne replied testily. “We’re working through a tasking issue, will advise shortly.”

    The two astronauts looked at each other. It was Holcomb who spoke first: “Rock-Paper-Scissors?”

    After a best-of-three, it was Ed who reluctantly twisted around in the cockpit to pull himself into the rover’s second Alice suit. If the cabin was beginning to smell like a sports locker room, the suits were closer to a pair of ski boots after a week in the Alps. The Orion hab included a supply of deodorant spray cans in its inventory, but for some reason NASA and its partner agencies had failed to include these in the rovers. The cans weren’t rated for transfer through vacuum, so they hadn’t brought them across before leaving. Ed was planning to raise this issue prominently in the post-mission “Lessons Learned” debriefings. Given his time again, he would have smuggled one over inside his suit, regulations be damned!

    Doing his best to ignore the odour, Ed methodically ran through the suit checks. This would be, what, his eighth EVA of the trip? By this point Ed was sure he’d be able to recite the checklist letter-perfect from memory fifty years from now, but the careful attention to detail drummed into all astronauts meant that he used the hardcopy on his wrist and took his time to make sure every point was covered. Checks completed, he disengaged from the suitlock and stepped off of the rover’s small platform and onto the lunar regolith.

    The view was familiar after three days of driving, but still stunning. The rover was perched on the rim of Shackleton, currently in sunlight but not in one of the Peaks of Eternal Light such as like the one hosting Orion. Looking down slope, Ed could see for a hundred meters or so before the crater plunged into darkness like a shore disappearing into the ocean. Even with his sun filter up, no details were visible in that inky pool. To the left and right he could see the occasional outcrop of rock as sections of the rim breached the terminator, but directly ahead there was just the pitch-black curve of the horizon blotting out the stars.


    But there was work to be done. Reaching up to activate his helmet lamp, Ed began headed down into the darkness with bounding strides. Over the past three weeks each member of the Orion team had settled on their own preferred means of extravehicular locomotion. Holcomb was a bunny-hopper, but Ed found he preferred this gentle lope, springing from one foot to the other. He found he could build up a surprisingly rapid pace if all he needed was to go in a straight line, as now, and it only took a couple of minutes to reach the edge of the sunlit region. It was only as he approached the shadow zone and started pushing back with each step to slow down that the full inertia of his body plus the suit made itself felt. Also making themselves felt were the blisters he’d earned on the ball of each foot from making exactly this maneuver over the past few days.

    Wincing slightly at the pain, Ed radioed back to Holcomb in the rover. “I’m at the edge of the shadow now.”

    “Roger that, I’ve got visual on you,” came her reply. “The map says the permanent shadow zone is about fifty meters further down, a little to your right. There should be a crater about twenty meters across almost directly ahead of you. If you skim the right side of that and keep going, you should get there.”

    Ed moved cautiously into the shadow, swinging his helmet light around slowly as he advanced, trying to spot a landmark in its dim puddle of light.

    “I don’t see... Ah, there it is! Okay, bearing right.” Now sure of his direction, Ed set off once more, this time keeping to a pace slow enough to be sure of spotting any potential trip-hazards. Fortunately, there seemed to be few rocks in this area larger than his fist, and the gentle slope was at a reasonably constant grade. Not much risk of making a Little Step here, Ed thought to himself, but better safe than sorry.

    “Okay, you can stop about there, Ed,” Holcomb told him. “You’re about at the right spot. How’s it looking?”

    “Unremarkable,” Ed responded, looking around. “The slope’s shallow enough for LITT. Should be no problem setting up the tripods.” He kicked experimentally at the surface. “Regolith is moderately thick, about four, five centimeters... Oh wow!”

    “What’s up, Ed?”

    Ed looked down at the shallow trench he had scuffed in the dirt, rocking slightly to change the angle of his lamp. Was that sparkle..?

    “I think we’ve got ice here! Just a few grains, but very close to the surface. I’m going to grab a sample.”

    The long handled scoop was back at the rover, so kneeling down in the articulated suit, Ed dug his gloves into the regolith and grabbed two big handfuls of the dirt. A couple of sample bags were still attached to the waist of his ALES, so he dropped both handfuls into one and popped the seal closed.

    “What do you think,” Holcomb asked. “Is there more here than over at Bussey Wells?”

    “Could be,” said Ed, scuffing his way around the site. “It’s certainly closer to the surface, and seems to be all over this area. Only four klicks out from the hab, too. Looks like this area could have more value than simply a place to put a telescope.”

    “Hold on,” said Holcomb, “I’ll bring the rover in closer and join you. We should get some more samples before we start calling this place ‘Boxall’s Brook’ or something.”

    Ed mulled that over for a few moments, before remembering how he had knelt down in the ALES to get his first sample. That wouldn’t have been possible in the old A9L suits. Maybe that should be commemorated somehow.

    “You know, I think I’d prefer the name ‘Alice Springs’...”



    Mission Day 35: From Russia with Love

    Yesterday Phil and I went for a drive in one of the open buggies left by Artemis 9 to see our Russian visitor, Luna-Pe. This was a much shorter journey than my recent traverse with Anne, so we took one of the open buggies left by Artemis 9, which has a larger trailer than the pressurised rover. Traffic wasn’t too bad, and the scenery was amazing. I took plenty of tourist snaps on the way, but unlike my previous surface excursions this wasn’t a science field-trip, but rather a supply run.

    Although the habitat and our partner cargo lander contain everything we need to stay alive on this first Orion expedition, future crews plan to stay a lot longer, and they’ll need a way to keep stocked up with essentials from home over the course of their mission. Of course NASA could handle that by just sending another full-sized cargo lander, but those big warehouses are expensive and a bit oversized for regular runs or last-minute replacements. Fortunately, our Russian partners have stepped up and agreed to help out by sending mail-runs on their Luna-Pe landers, the first of which touched down last week just a couple of kilometres from the base.

    When Phil and I got out there, we found the lander sitting happily behind a low rise, with a pile of goodies stacked on its back waiting for us. Whilst Phil backed the rover up to the base of the lander, I climbed a ladder on Luna-Pe’s side and hooked up the first cargo box to Pe’s little crane. This was a bit tricky in my space gloves, but the Russians had made their controls nice and chunky so even an uncoordinated space monkey like me didn’t have too much trouble.

    We had the rover loaded up in less than an hour, then made the twenty-minute drive back to Orion before emptying the trailer and heading back for the second load. That second run went even smoother than the first. It was much easier to hook up the remaining crates with less stuff cluttering up Pe’s cargo deck.

    Even though we were quicker on the second run, we were both tired by the time we got back to base, so Houston agreed we should hold off unpacking our goodies until today. This morning, Anne and Winch went out and opened up the first crate, which included a new set of “Crew Personal Preference Kits” (NASA-speak for care package from home). Once they’d brought them into the airlock (along with the inevitable swarm of dust - have I mentioned the dust?!), we opened them up like kids on Christmas morning.

    Compared to six-month Freedom expeditions, our little six-week camping trip might seem not so long, but it’s still nice to get reminders of home. My CPPK included a memory stick of the new Star Trek movie and an “I Heart The Moon” T-Shirt, but the best thing I found inside was a supply of those little airline containers of pasteurised milk and teabags. No more weapons-grade coffee for me in the mornings! Thanks to our Russian postman I can now enjoy a proper cup of English Breakfast tea with milk.


    “Jeez, I’d almost forgotten how awkward these things are to put on,” Chung groused from the airlock as he fought to lock the pants of his A9L moonsuit to the torso. “Hey Phil, you sure we can’t take the Turtles with us?”

    “Forget it, Winch,” Commander Whitt replied. “This is a base now, not a sortie outpost. We can’t tell the next crew they have to stick to their A9ls just because you’re having trouble fitting your beer gut back into your old suit.”

    Ed listened to the banter with half an ear as he finished clearing his bunk area in Orion’s dome. No, not his area, not anymore. As Phil had pointed out, the small metal and fabric hut they’d called home for the past month would soon be left empty, waiting to host a new crew. Someone else would be sleeping here next year, and probably someone else again the year after. This was never going to be more than a temporary home for Ed, no different really from the endless anonymous hotel rooms he’d used over his years of training.

    No, of course that wasn’t true. Orion was more than a place to sleep. It was their protector and comforter in a barren, hostile, beautiful land. Outside of this Hab and the Clarke, there was no-where else on this entire world where a human being could survive and thrive. The Apollo pioneers had proven the Moon could be reached, whilst the first Artemis sorties had shown how it could be explored. With Orion, humans had finally demonstrated that they could settle down and live on this, the rocky shore of the interplanetary ocean. Whilst Orion was crewed, humanity had two homes in the solar system.

    The crews that followed would build on that legacy, extending their stays until finally settling permanently on the Moon. But for the crew of the first Orion expedition, it was time to return to the mother-world.
     
    Part IV, Post 18: The Planetary Science Community Strikes Back
  • Hello everyone! Sorry about this being up a bit late tonight, but it's finally that time again. Last week, we were very pleased to bring you Nixonshead's guest post about the crew of the first human outpost on the lunar surface. However, this week, we're turning the attention to the fight on behalf of the US unmanned program in 2005, when the lack of a new Pioneer selection made it seem like the end...

    Eyes Turned Skywards, Part IV: Post #18

    Although planetary scientists had been moved to depression and anger by NASA’s apparent abandonment of the Pioneer Program at the beginning of 2006, the fact that most were college faculty or students beholden to teaching and class schedules meant that it was not until early June that a group of several dozen of the field’s most eminent scientists were able to heeding the call of Cornell’s chair of astronomy, Jay Lawrence, and descend on Ithaca, New York, for an informal meeting on the future of the field. Although the over four months that had passed since his invitation had allowed tempers and passions to cool, they had also had the effect of allowing the invitees, all of whom had strong reputations for honesty, fair dealing, and scientific rigor, to begin hashing out the beginnings of a plan even before the meeting itself. As expected, all agreed that planetary science deserved better than what it had gotten, but, at least at the beginning of this process, they were unable to agree on much else. Everything from scientific priorities to destinations to individual missions was up in the air. Flurries of emails sallied forth from departments around the country to do battle, attempting to make the case that what project their writers favored was, in fact, the best path forwards for American planetary science.

    Fortunately, Lawrence and his friend and co-conspirator Jonathan Mills, a senior faculty member within Columbia University’s planetary science group, were able to soothe tempers and keep the focus of the invitees firmly on the external challenges faced by planetary science, without adding internal conflict to the agenda. While it might seem of vital importance to decide whether to venture towards the ice giants, Venus, Mars, or Jupiter, they emphasized that to an outsider--a Congressional representative, say, or a member of the press--such disputes would seem arcane and technical, and reduce their willingness to support any planetary program. Matters were no doubt simplified by the fact that the invitees had specifically been chosen based on their reputation for supporting the best missions, not just the best for them, and perhaps assisted by Lawrence’s constant references to astronomers, who had been far more successful in obtaining a steady, active space program than planetary scientists despite facing similar questions of utility and cost.

    By the time the entire group was sitting down face-to-face in a hired conference room in Ithaca (pointedly not Cornell, to avoid the appearance of university support), the broad outlines of what would become known as the “Cornell Plan” had already been laid. Something of a masterpiece at being all things to all people, the Cornell Plan eschewed specific mission details in favor of outlining a mission-planning process, allowing any planetary scientist to imagine his or her preferred missions getting the nod, and any politician to imagine steady, predictable, and low budgets. Besides finalizing the slim booklet they would publish on the plan, the meeting focused on developing the public-relations operation they would use to sell it, allowing them to develop a strategy for advocacy over a week of dense discussions. While they could publish statements all they wanted, they were, in the end, asking for billions of dollars of money from the government of the United States, and would need to sell the planetary exploration program, not merely politely request that it be expanded. The first salvo in the public relations offensive they were about to embark on was prepared before they left Ithaca, in the form of an open letter they had sent to the New York Times, the Washington Post, the Houston Chronicle, and several other national or particularly space-interested newspapers.

    Drafted largely by Lawrence and Mills, the letter began by acknowledging that the astronauts of the Artemis missions had greatly advanced the state of the art of lunar science, but then quickly pivoted to point out
    no astronaut will step out of a capsule to stir the sands of Venus, nor dare Jupiter’s might to swim Europa’ seas, not in this century. To search out, explore, and discover these other places, these realms beyond human touch, we must send machines, robots, in our stead, to brave the dangers we cannot risk.​
    The letter then continued with a lengthy justification of the value of robotic explorers and planetary exploration in general, both through practical examples such as Venus’ role in climate change research on Earth and through more soaring rhetoric claiming a need and desire to explore the solar system. However, it noted, aside from the Pioneer program and Artemis missions, there had not been a single major planetary probe approved for development in the nearly twenty years since Cassini. Worse, with the recent cancellation of the 2006 Pioneer selection it appeared that far from continuing this highly successful scientific program the Administration was on the verge of completely abandoning robotic space exploration, leaving it to Europe, Japan, and other nations to carry human eyes and ears beyond the orbit of the Moon. Only action now, it concluded, could hope to maintain America’s place at the forefront of planetary exploration.

    As with the Cornell Plan (which the letter was an even further simplified version of), however, Lawrence and Mills knew that one letter, even if it was widely published, would not suddenly spur Congress to allocate hundreds of millions more dollars for planetary exploration. For their second salvo, therefore, they turned towards a more potent lever by visiting the Washington offices of the National Space Organization. While the Organization’s break-up with Carl Sagan in the late 1980s had not been entirely amicable, it still retained substantial imprints from his directorship, particularly a (theoretically) strong commitment towards promoting robotic spaceflight. More importantly, it also had the largest membership of any space advocacy organization, with over 50,000 dues-paying members--less than half its peak in the early 1980s, but still a potentially formidable force if mobilized properly. By forcefully arguing the value of the planetary exploration program to space exploration and development more generally, they were able to persuade the Organization’s leadership to back their campaign to revive it. Soon enough, emails began landing in the inboxes of the Organization’s membership, urging them to call or write their Congressional membership. Many did, and though the Organization’s membership was too scattered to make a dent in the Congressional mail system, it certainly had effects on members from more space-interested districts like those containing NASA centers.

    To wrap up their public relations offensive, the scientists turned at last towards their main target, Congress itself. As ever, the main enemy space advocates had to confront was indifference, not outright dislike. Most members of Congress, as with most Americans, liked robotic explorers when they turned up on the news; they appreciated the images sent home by probes such as Galileo or Cassini, or the drama which had surrounded the landing of the Mars Traverse Rovers and the inability to free the trapped “Liberty” rover. They even wondered at the scientific results from these spacecraft. And they felt that the United States should continue doing such things, as much out of a vague sense of wanting to push the frontier as anything else. What they did not think about, or perhaps know about in some cases, was what was being spent on those spacecraft, nor what was needed to keep the American planetary exploration program, nor, even, where that money was being spent. Most members of Congress, after all, did not have NASA field centers in their districts, and even with NASA’s mastery of contract distribution to all corners of the nation, few of them would obviously be affected by a further contraction of the planetary science program.

    Therefore, the goal of the scientists was, primarily, to educate Congress, to make them aware of what the new budget meant and what planetary science meant for their districts, in both financial and non-financial senses. Many recruited their graduate students, staff, colleagues, and contractors to call, mail, or email their local Congressional representative to explain the amount of money that NASA’s programs brought to their district, often sums in the millions of dollars, and how many people they employed, both directly and indirectly, while also pointing out that the current planetary science program and its great successes had only been created by investments in the 1980s and 1990s, which were now on the verge of drying up. Others reached out to their local communities, whether via the media or through appearances at museums, planetariums, or simple community events, spreading knowledge about the planets and about how more funding was needed to continue the series of planetary probes that had been running since the 1970s, especially to new, exciting destinations like the undersea oceans of Europa or Enceladus, or to old destinations that had shown signs of needing a second look, like the curiously methanated plains of Mars.

    All of these measures, naturally, had more influence in districts where the planetary science program and space exploration more generally were already important. In Houston, Los Angeles, San Francisco, Orlando, and Baltimore, politicians were already primed to consider space exploration an important issue worthy of action. Many of the areas where the academic programs were largest, however, had never thought of themselves as having much direct connection to the space program, something these scientists were beginning to disprove. A more vigorous planetary science program, their representatives were learning, could bring federal dollars to their area--always an attraction--and, moreover, do so without any appearance of logrolling, favors trading, or, in short, pork at all, if they adopted the Cornell Plan’s recommendations. Thus, these representatives slowly began moving to investigate planetary exploration as a subject for Congressional action, as the Cornell Plan’s authors were only too happy to help them do. Their most immediate goal was to ensure that the Pioneer Program, at least, continued, something that could be done easily enough and which promised quick and easy action. As summer began to draw to a close, a series of hearings on the issue were begun on Capitol Hill to investigate just why NASA had chosen to stop the apparently successful Pioneer program.

    The entire affair, it turned out several months later, had ultimately boiled down to apathy compounded with poor communication by NASA administration. The Woods budget had trimmed funding for the planetary science program, including the Pioneer budget line, something no one in Congress had caught, and Headquarters staff had determined that a Pioneer selection that fiscal year would not be feasible. However, a slight delay, to the next fiscal year, could permit work to carry on almost as usual, so they had decided to push back the selection announcement until September. Unfortunately, they had also decided to remain quiet until the new selection announcement could be made instead of mentioning this fact, exploiting ambiguities in the rules set up for the Pioneer program to avoid embarrassment for the agency and the President. Of course, this had now backfired, and the offending staff were reassigned or dismissed, while Congress ordered the agency to make a selection as soon as possible. Even before they did so, pro-space representatives had managed to insert increased Pioneer funding into the FY 2008 budget and added language requiring the agency to make a selection every other year, “beginning with the next normal selection in FY 2010.”

    While Congress was investigating the Pioneer program, the scientists who had pushed it into doing so in the first place were still working, this time to convince Congress to institute a larger overhaul of how NASA selected planetary science missions, especially ones beyond the capacity of the Pioneer program. After all, as they had been saying from the beginning, there were plenty of missions that were simply beyond the budget and schedule guidelines of Pioneer, but at the same time the haphazard, arbitrary, and politically-driven mission selection process that had operated since the 1960s had clearly fallen apart in the face of greater budgetary challenges and diminishing top-level attention. While simply copying the Pioneer process was untenable due to the sheer expense of Cornerstone-class planetary science missions, a more science-driven approach, similar to the Pioneer process, or to project selection processes in many other fields of science, such as astronomy, nuclear physics, or particle physics, was greatly tempting. A similar planetary science process would separate the debate about which missions were most scientifically viable from the question of which missions would be launched. At the same time, doing so would provide Congress with a tool to check any wasteful spending, ensuring that NASA would always be carrying out the most scientifically valuable missions instead of the best at, say, lining the pockets of the President’s friends. Although it took longer than rectifying the Pioneer situation, this, too, became enshrined in law in early 2008. After another year of organizing, the newly establish Planetary Science Prioritization Panel, or PSP^2, began meeting in early 2009 to begin drawing up an outline of the next decade of American planetary science missions, eager to return NASA to the forefront of planetary exploration.
     
    Part IV, Post 19: Reusable space vehicles
  • Good evening, everyone! I hope you'll find a thing or two to enjoy in this week's Eyes post...

    Eyes Turned Skyward, Part IV: Post #19

    Back in 2002 when competitors had contemplated Thunderbolt and the potential value of reusable vehicles, they had presumed the benefit of time to evaluate the still-immature technologies involved and plan their own reactions. However, the announcement of the Northrop TransOrbital division the next year had been a serious blow to this complacency, and studies throughout the US and around the globe had acquired a new degree of urgency. By 2006, almost four years later, the preferability of developing their own reusable vehicles was no longer up for serious debate--with more than 50 launches having cut the skies of the Northeast (launches from Wallops were commonly visible to tourists on the Washington Mall, and occasionally from New York City), the age of the reusable vehicle had clearly arrived. Moreover, the technology TransOrbital would use to achieve its “trans-shipment” of payloads from solely-LEO-capable light launchers to geosynchronous transfer orbits saw a serious risk reduction with the only-slightly-delayed launch of NASA’s Cryogenic Depot Demonstrator on a Delta 5060. Though the CDC alone was in theory within Thunderbolt’s capacity, the launcher wasn’t yet certified by NASA for such high-value payloads, and the increased launch capacity of the 5060 meant slightly more of the depot’s total theoretical 40-ton propellant capacity could be loaded, providing a better test of the depot concept.

    Despite the high hopes resting on it from NASA, space fans, and Northrop’s management, the CDC’s operations were utterly routine--the spacecraft deployed its solar arrays, shifted residual ascent hydrogen from the converted Centaur second stage into the orbital storage module, purged the tank, then pumped the liquid oxygen residuals from their own tank into the now-empty hydrogen tank. With the main business done, there was little for controllers on the ground to do but wait as time passed and the effectiveness of the demonstrator at retaining oxygen and particularly hydrogen became clear. With Northrop engineers aiming to design a fully-operational sister depot for TransOrbital chomping at the bit for data, the months it took for full data to come in were an anticlimax: an exercise in patience and frustration. Finally, the initial results came out: in its first three months in orbit, the initial propellant load of eight metric tons had fallen to just over seven and a half. This loss of about 0.07% per day was concentrated mostly in the depot’s hydrogen supply as the station allowed hydrogen to boil off, serving for both a heatsink for the liquid oxygen tank and for orbital station-keeping.

    Though for Northrop’s purposes this result was quite sufficient, the loss rate posed a problem for NASA--Northrop’s TransOrbital depot would be topped off potentially once a month or more, and its capacity “turned over” several times a year, while NASA was curious about storage at more distant destinations like EML-2 or on board spacecraft traveling to and from Mars--destinations which might require storage duration measured in years, not months, and where orbital station-keeping wasn’t a major need. NASA therefore began revisiting at low intensity the potential to incorporate active cryo-cooling into follow-on depots, to open up placement and supply of such distant depots. For their part, Northrop turned their attention fully to testing their tug architecture, rolling information from the CDC into their operational depot, and (most importantly) selling the service to customers. This process was helped by the ongoing progress in technology risk reduction , such as the mid-2007 launch of their first tug and modified Centaur “dumb tanker” on a Northrop Thunderbolt, which saw the tug separate, maneuver on its internal propellant, return to dock and pump fuel from the second stage, then alter its own orbit several times. However, in addition to these achievements and the lure of a near-50% cost reductions over their ELV competition, TransOrbital also found a new ally in selling the TransOrbital concept: Lockheed-McDonnell.

    Despite operating their own X-33 Starclipper demonstrator before Thunderbolt’s debut tests, Lockheed had (in the view of some) squandered their chance to make lemonade out of the failure to pluck the lemon of reusable single-stage-to-orbit. However, internally and in great secrecy, Lockheed had in fact been using the example of the Thunderbolt--and their own Starclipper knowledge and team--to pursue risk abatement and studies on their own potential RLV. Indeed, some within the team had been advocating for an immediate announcement before the arrival of TransOrbital had shaken the commercial scene, stymied mostly by the sheer scale of a vehicle required to match even Delta’s payload to geosynchronous transfer orbit, much less their new European and Russian competition, as well as the thermodynamic challenges introduced by returning from GTO compared to a low Earth orbit. The implications of Northrop’s announcement had much the same effect on this continued Starclipper project as it had on Thunderbolt--it put a commercially competitive launcher in reach with a smaller overall vehicle by dividing the mass required in LEO over multiple launches and enabling the launcher itself to be focused on just reaching and returning from low orbit.

    Thus, as satellite builders themselves, Lockheed management had co-operated readily with TransOrbital’s design pitches on creating “TransOrbital-capable” variants of various popular comsats and adopted a wait-and-see approach while Starclipper development was internally focused on a fully-reusable Thunderbolt-class vehicle. In 2007, Lockheed judged the time right--sufficient risk had been eliminated that it seemed safe to bet on TransOrbital reaching operations. Thus, they publicly unveiled the “new” Starclipper concept, which called for a two-stage, fully-reusable launcher. The first stage would be a scaled-up refinement of the familiar X-33 design with roughly double the propellant load of the earlier vehicle. This booster would launch from the Matagorda Spaceport, which American Launch Services had been glad, this time, to share with a new tenant, having turned down a similar operation-sharing offer from Thunderbolt before StarLaunch had proceeded to absorb much of their business. Unlike Thunderbolt, which needed to retain substantial propellant in the first stage to boost back to the launch site, Starclipper’s first stage would instead contribute as much delta-v to the ascent as possible, building its own downrange velocity to enable it to glide forward to a new landing site on the Gulf Coast of Florida. From this site, where the booster would land like an airplane, the vehicle could either be barged back to Matagorda or (if the regulatory environment was more favorable) refueled and flown back on its own power.

    For the orbiter, Lockheed wanted to ensure sufficient payload volume for geosynch-bound satellites, and thus planned a much larger cargo bay than would normally be included given the system’s targeted 10-ton payload capacity (aimed to enable them to launch the entire Delta national security line, as well as some of the smaller M02-class payloads which had been launched without competition on Multibody for decades). This large bay--5 meters across by 8 meters long-- and relatively small fuel load drove the Starclipper team away from the X-33 wedge shape, instead selecting a design closer to the delta-winged fuselage of the X-20 Dynasoar, the Japanese HOPE-C, or the long-dead Space Shuttle concepts. More like a craft out of science fiction, the propellant tanks for the orbiter would make up only half the orbiter’s fuselage, with the LH2 tank placed forward of the bay and the LOX tank aft. In the tail, a set of pod-mounted hypergolic thrusters would bracket the single J-2S main engine to form the Orbital Maneuvering System, which would be used for any orbital alterations or corrections as well as for initiating the vehicle’s descent. Several landing sites were proposed in the rollout, ranging from the same gulf-coast Florida site as booster recoveries to California to landings at Cape Canaveral itself, with airlift or self-ferry back to the launch site.

    The proposal was a mix of technologies from the long-proven (such as the orbiter’s engines and both stage’s automated flight modes) to the tested-but-new (the metallic TPS tested on X-33 and fitted to both booster and orbiter, the booster’s altitude-compensating aerospike, and flight-proven lobed aluminum tanks) to the novel (the orbiter’s use of composite structures for the less-complex cylindrical propellant tanks and primary structure components), and the proposed results were striking--not only would the Starclipper beat Thunderbolt’s $2,500/kg price point, it’d be available for just half that. Moreover, based on their five years of secret internal development and refinement, Lockheed announced a goal for powered test flights within five years, with operational service to follow “shortly after.” The concept attracted immediate attention, not the least from StarLaunch, who had been experiencing difficulty funding their own reusable L2 second stage. Other early interest came from TransOrbital, who were intrigued by a backup--and potentially even cheaper--propellant provider. Lockheed, for their part, had prepared several images of a Starclipper carrying a propellant tank in its bay to such a depot, as well as trans-shipping payloads via TransOrbital. NASA, too, expressed some interest--with Starclipper able to maneuver itself, it could potentially carry a cargo pod similar to the Apollo Mission Module to dock to Freedom or a successor station with a substantial payload of supplies, even if crew-rating the vehicle might prove challenging--a potential way to supplement or replace the Aardvark logistics vehicle with a dramatically cheaper option.

    Lockheed, however, wasn’t the only group considering their own reusable launch vehicle. The European space community had been pursuing reusable launchers since the 90s, stifled first by the failures of the Sanger II’s turborocket systems, then the political infighting and constrained budgets which had surrounded the RLV question in the aftermath of the introduction of Thunderbolt, compounded by the 2004 Recession. However, Thunderbolt and the new impetus of Starclipper were enough to finally set the gears of the bureaucracy binding the British, French, German, and Italian space programs together into motion. Beginning in 2007, talks between European space ministers and executives of the Europaspace consortium picked up, focused on a unified European response to retain the dominance in commercial space launch Europe had seemed poised for in the late 90s with the retirement of the American Titan and the introduction of the cheap and capable Europa 5. These talks dragged on, complicated by the need to apportion work appropriately between participating nations and the complexities of funding split between ESA and Europaspace, but finally in 2008 a plan emerged that all parties could be satisfied with, one which should see Europe matching American reuse capacities by the middle of the 2010s, as well as putting them at the forefront of future developments in reuse.

    As with Starclipper, the primary goal of the 2008 reusability plan was the development of Aquila, a fully-reusable two-stage launch vehicle sized for a capacity of roughy 30 tons to orbit--having finally achieved the lofty payload possible with the American Saturn and Russian Vulkan, Europe could hardly take a step back in capacity. The orbiter, Ganymede, would be based on the Horus spaceplane that had seen glider tests almost a decade before, using a new French-developed hydrogen/oxygen gas generator engine. As Horus itself had been German-led, so too would the Ganymede operational variant. Ironically, unlike with Starclipper, with the European plan it was the orbiter vehicle which was closer to previously developed and tested hardware, with the main challenges anticipated for the Horus spaceplane being the scaling up of the vehicle.

    To start Ganymede on its way to orbit, the French and British members of the consortium would combine forces on the new booster stage, Aetos. Here, the Europeans were faced with problems of geography: unlike Lockheed’s planned operations from Matagorda, French Guiana had been specifically selected for having no troublesome land for thousands of kilometers downrange, ensuring that any first stage would have to return to Kourou for reuse. This narrowed the reuse options to a rocket-powered first stage, similar to a dramatically scaled-up Thunderbolt first stage, or a winged booster fitted with jet engines. Though the systems were comparable in weight, internal development at Europaspace focused primarily on the winged flyback booster, since once ignited, the jet engines would give the booster sufficient range to recover from vast distances downrange, and thus minimize the need for a lofted trajectory to reduce downrange distance, boosting achievable first stage contribution to the ascent. There were also both cultural and organizational reasons: designing a winged vehicle would draw on the extensive experience of European aeronautical firms (such as those making up Airbus, another part of the same European conglomerate which owned the private shares of Europaspace), and assist in spreading development funding and responsibilities within Europe. The fact that this would result in Aetos, named for the Eagle of Zeus, bearing more resemblance to the concepts of reusable vehicles which had permeated aerospace since the era of Von Braun was also valuable, and not entirely a coincidence--when asked for a reusable vehicle, many still pictured the reusable spaceplanes of the 50s, 60s, and early 70s.

    The second leg of European development efforts would be focused matching this reusable launch system with new developments in orbital operations to enhance its capabilities. The Italian-led effort focused on the newly-important “orbit-to-orbit” area of space missions, with the goal of developing a new, all-European version of the American TransOrbital system, pairing a reusable orbital tug with LEO depots. However, unlike the Centaur heritage of Northrop’s design which meant its capacity was limited by the need to carry propellant to brake the tug back into low Earth orbit to refuel and pick up new payloads, the Italians planned to start from a clean sheet and aim for maximum capability by making use of Minotaur experience with ballistic entry to design a tug which would “aerocapture” back to LEO from higher energy orbits like GTO or even (potentially) lunar return trajectories, scrubbing off velocity in the rarified upper atmosphere protected by a (relatively light) reusable thermal protection system. Once operational, this Phoenix tug would thus be more capable and fuel-efficient than than the TransOrbital Centaurs, and the only cost would be higher development costs. However, with recovery from the sharp but short 2004 recession ongoing, investing money in supporting the European industrial base wasn’t unwelcome, and in fact constituted something of a benefit.

    The final element of the European plan was aimed at the future, ensuring that the European community would end up out in front in any further evolutions of the reusability field. For almost five years, a small team of former Rolls-Royce engineers lead by Alan Bond had been pursuing a new design for a combined-cycle air-breathing rocket engine which they had originally developed in the 1990s. However, between their departure from the main British space establishment and 2008, the team had made what they believed to be several key breakthroughs, mostly derived from a new precooler concept and the insight of not seeking to fully liquify incoming air. Combined with a reconfigured vehicle layout, they believed that a fully reusable SSTO with payload comparable to the emerging TSTO RLVs might be achievable before 2025. With the overall spending on reusability efforts in Europe picking up, the project now finally found interested ears, particularly within the British portion of the European space development community. The team thus picked up several contracts for low-cost efforts aimed to validate the engine and precooler’s design assumptions, and confirm if the system were at all feasible. If successful, the system could be pursued as a second-generation reusable vehicle to follow-up on Aquila.

    However, while their competitors were beginning work on their own full-scale reusable vehicles, engineers at Star Launch Services, who had begun the reusability boom with their Thunderbolt vehicle, were hardly idle. After all, Thunderbolt had always been conceptualized as part of a fully-reusable vehicle, not dissimilar to the European and Lockheed systems. However, unlike those firms, Star Launch had no major government ties or alternate revenue to draw upon, and had to fund L2 development entirely from what they could afford internally or draw from outside investment. Thus, despite the head start of operating Thunderbolt for more than 6 years, the L2 design was still in a relatively early design stage. The core concept was an adaption of the Thunderbolt first stage with a payload bay sandwiched between the five meter diameter LOX and LH2 tanks. However, unlike Starclipper and Horus, Thunderbolt’s second stage would re-enter tail-first. Several options were considered for this heat shield and the incorporation of the vehicle’s main engines. These included a ceramic tile system including doors closing over conventional bell nozzles, a regeneratively-cooled annular “plug” aerospike which could serve as combination of engine and heat shield, and more novel systems such as a “transpiration” setup which would flow residual cryogenic hydrogen into the base plasma flow to absorb and carry away heat before it could reach the vehicle. As Allen and Hunt’s team worked to address these questions, the firm’s lead in reusable vehicles was eroding. They would be lucky to get the L2 funded, designed, and tested by the time that Lockheed and Europe were planning to introduce their own vehicles.

    The question of developing reusable vehicles had even reached the halls of NASA itself. Excitement over the achievements of Thunderbolt had long circulated, as well as debates over redeveloping the Saturn system to be capable of similar feats, though with substantially larger payload capacity However, with the pressure of Freedom and Orion support, as well as ongoing development of their respective successors, NASA was--for all its budget--in less of a position to begin a complex development program, and the concepts remained confined primarily to design and operations studies. Like the Europeans at Kourou, NASA’s launch site at Cape Canaveral was specifically sited to avoid overflight of land downrange, which now became a challenge in potential return and reuse of some sort of Saturn-scale system. However, the payload hit of designing a full boost-back system like Thunderbolt was steep--as much as 30%, for first stage reuse alone. Given the specific goal of retaining Saturn-class payload capacity, this was unacceptable. However, unfortunately, the advantage of a winged, jet-powered stage like the European plans were relatively minor, as the weight gain from wings and turbines would be similar to that required to boost back to the launch site on the stage’s own rocket engines, and require much more dramatic departure from the existing Saturn design.

    In the end, NASA’s research contracts with Boeing focused on a compromise approach: Saturn first stage tankage placed atop a thrust structure modified with landing gear and terminal descent engines flanking the existing F-1A. Like Saturn, this new core would be capable of being clustered into a tri-core “Heavy” configuration. For flights where the payload capacity could be spared, the first stage would conduct a full return to launch site as with Thunderbolt, and some research was begun to determine the feasibility of cross-feeding propellant from the boosters into the center core to deplete the boosters faster and separate slower and closer to land. However, for flights where payload capacity was critical, the core would instead conduct a recovery downrange on a new landing barge, which would cut the “reuse penalty” for the stage almost in half. The second stage which was intended to be paired with this first stage was much more notional, ranging from L2-style near-capsules to winged orbiters like Starclipper and Horus, and in fact both concepts were studied in “Phase A” contracts commissioned from several firms. However, given the cost benefits for reuse were found primarily in the first stage, NASA placed priority on determining the feasibility and operational concepts of this new, reusable Saturn core, in the hopes of seeing it approved and development begun. Considering the other projects on their plate and the re-election of the budget hawk President Woods, it would be an uphill battle to achieve even that.

    While plans for reusable systems around the globe were percolating, at least one system was reaching milestones in its introduction. In early 2008, a Lockheed Delta carried the Transorbital depot to orbit. There, it underwent a series of storage and commissioning trials over the next several months, similar to the ones conducted two years earlier on its near-sister, the NASA CDC. However, August saw the differences between the NASA demonstrator and its operational derivative put to the test when the tug launched the previous summer met a Thunderbolt tanker launch, filled its tanks from the Centaur second stage, then carried the fuel to the depot to begin the process of topping off the station’s tanks. The transfer worked just as had been planned, and the system was ready for its first trial run. As Thanksgiving approached, a Lockheed-built mass simulator/demonstration payload was prepared and launched on another Thunderbolt to meet the tug, which had already undocked from the depot and been positioned for pickup. The tug successfully locked onto the payload’s beacon, and ground controllers put it through its paces to phase its orbit to match the payload as it had with the tanker flight in August.

    The two made rendezvous the day after launch, and after a careful docking to the simulated satellite’s CADS port, the tug lit its engines to push the combined stack into the intended GTO and separated from the “satellite.” Once the maneuver’s success was confirmed and the achieved orbital accuracy plotted, the tug redocked to the satellite and made another short burn to drop the stack’s perigee into the atmosphere before separating a second time--a maneuver aimed to both further demonstrate docking procedures as well as minimize orbital debris [1]. The tug, alone once more, then fired its engines to raise its own apogee to safety, coasted back to low Earth orbit, and burned again to drop into LEO and return to the depot for refueling. The gamble had worked, and TransOrbital Services was open for business none too soon, with their first commercial customers expecting launches in the summer of 2009. It would take many more missions before the process would become routine and considered totally risk-free, but the cost benefits were tempting. In the early 2000s, the benefits and risks of reusable systems had been up for debate, but with the ongoing commercial success of Thunderbolt and the technical achievements of TransOrbital, these systems were clearly only the beginning. The end goal wasn’t the partial reuse of these first-generation systems, but full reuse, and the race to achieve it was on as StarLaunch, Lockheed, and Europe vied to achieve this next stage in reusability development...

    [1] A trajectory similar to EFT-1 IOTL
     
    Part IV, Post 20: Space based astronomy in the 21st century
  • Good evening, everyone! I actually remembered the implications of it being Friday after the gym today, so this week's post comes to you on time, for once. :) Last week, we talked about the revolution in reusable access to space, but this week we're looking at some of the payloads that enables: our eyes on the sky, in the sky, space telescopes.

    Eyes Turned Skyward, Part IV: Post #20

    With the launch of the Compton Gamma-Ray Observatory in 2005, high-energy astronomy entered a new and distinctly different phase from anything that had come before. Never before had so many instruments of such power flown simultaneously. In addition to Compton itself, equipped to see gamma-ray photons generated by only the most energetic cosmic events, there was Leavitt, able to peer into the lower (though still ferociously energetic) x-ray bands, the Particle Astrophysics Magnetic Facility (Astromag) on Freedom, sifting through the particle detritus left over from the same processes that had produced the light seen by Compton and Leavitt, and a host of optical, infrared, and even radio telescopes, both on Earth and off of it, poised to bring an unprecedented array of viewpoints to bear on the questions of the cosmos. In theory, if all of the instruments built around the world over the past several decades and used to study the heavens were used in concert, the activity of galaxies and stars could be examined from the mega-electronvolt output of their most energetic flares and paroxysms to the gentle radio hum of chemical activity slowly churning in vast, cold molecular clouds, spanning over twenty decades of spectrum, in the terms used by astronomers.

    In practice, of course, there were too many specialized research programs—too many scientists chasing too many hypotheses in too many countries—for such a massively integrated research program to appear. Instead, the observatories of the world clicked through their own research programs, with only the occasional joint production, as Compton and Leavitt looked towards the same targets, or one of them and one of the giant new ground-based telescopes beginning to sprout around the world. Nevertheless, these occasional period of cooperation proved to be tremendously productive, even if they had been anticipated by nearly a decade by a series of more specialized spacecraft, designed to observe on multiple frequencies simultaneously.

    The reason for building a generalized spacecraft like the High-Energy Event Explorers was simple: there was a problem only they could solve. Since the 1960s, with the launch of the Vela satellites by the United States Air Force in an effort to track nuclear testing worldwide, it had been known that, at irregular but frequent intervals, mysterious and powerful bursts of gamma-rays would appear, linger for a brief period of time and disappear in the background of space, with no detectable progenitor events waiting behind them to give a clue as to where these strange events could be coming from. Years of study and research through the 1970s and 1980s, once these events had been declassified, had yielded little progress, and the original plans for the Compton Gamma-Ray Observatory at that time had included a special instrument for studying these gamma-ray bursts. When it became clear that the Observatory would not even start construction for a decade or more, however, scientists interested in these odd gamma-ray bursts began to push for a specialized, dedicated mission, specifically designed to detect and map gamma-ray bursts and determine once and for all their distribution and origins.

    The result was the first High-Energy Event Explorer. Naturally, it had a sensitive gamma-ray detector aboard as its first and most important instrument; gamma-rays were, after all, the signature of a gamma-ray burst. But it also had a small but capable x-ray telescope, intended to work together with the gamma-ray burst detector to find the cooler, longer-wavelength light that must be emitted from such a powerful event as its initial, prodigious energy output waned. While modest in size, it was designed to provide unprecedented precision in localizing any progenitor, far above and beyond the modest experiments that had been launched so far, and unprecedented swiftness in slewing these telescopes towards their targets, courtesy of on-board software capable of steering the spacecraft towards gamma-ray bursts without ground intervention. Over its six year mission, from 1989 to 1995, when the last of several reaction wheels needed to precisely steer the telescopes towards their targets failed, HEEE-1 created the first all-sky map of gamma-ray bursts, showing unambiguously that they were uniformly distributed on the sky. This seriously dented, though did not completely destroy, the hypothesis that gamma-ray bursts had a source near the Milky Way, instead suggesting a distant, extragalactic origin.

    More importantly, however, HEEE-1 identified the first known GRB progenitor when, in late 1991, it caught a GRB as it was first flaring, fast enough that the newly commissioned and incredibly sensitive Keck Telescope on Mauna Loa was able to slew towards the coordinates in an emergency observation, catching the first faint traces of a GRB’s optical afterglow. As the burst’s light faded, these observations allowed astronomers to determine its location to exquisite, unprecedented precision, pinpointing the burst’s origin galaxy. Follow-up measurements of that galaxy’s redshift showed that the light that HEEE-1 and Keck had detected had come from a galaxy that was far, far away, billions of light years, and had been emitted when the universe had been substantially younger. This observation, while working together with HEEE-1’s all-sky map to answer one question, had only raised another: if gamma-ray bursts were coming from outside our own galaxy, indeed from galaxies billions of light-years away, then a few basic calculations showed that the events that gave birth to them must have released a truly astronomical amount of energy, far outstripping even the most powerful types of supernova theoretically possible. What, then could be generating such events, if nothing much short of the Big Bang seemed adequate? To answer this question, astronomers needed more data—and another mission to collect it.

    This follow-up—the second High-Energy Event Explorer, or HEEE-2—would not be launched for nearly a decade, just beating the Compton observatory itself into space with a launch in 2004, but quickly began to prove its worth in conjunction with the Japanese Chasen observatory, another satellite specialized in gamma-ray burst observations. HEEE-2 took the basic design of the first HEEE and tweaked it, adding a small optical telescope to allow the spacecraft to track ultraviolet and visible emissions from located gamma-ray bursts without needing ground-based support and upgrading the spacecraft’s ability to precisely locate gamma-ray bursts and other transient events, allowing a new era of mechanized detection of bursts to begin. Over the past decade of operations, HETE-2 and Chasen have detected hundreds of gamma-ray bursts, which together with observations from other telescopes began to uncover the details behind their formation, showing them to most likely be the result of certain special types of supernova lining up in just the right way to aim powerful, luminescent “jets” at Earth, appearing far brighter than even their massive power would normally indicate.

    With on-going study clearly locating gamma-ray bursts in other, distant galaxies, the ever-growing coalition behind the Large Infrared Space Telescope found that it could count on another band of supporters. With Leavitt, at least, having many productive years ahead of it, and Compton having just started its observational career, many high-energy astronomers were now interested in having a new infrared telescope to complement their data, one that could peer back into the distant past and begin to uncover the earliest galaxies that were giving birth to these hyper-energetic events. This was part of the reason that LIST saw such dedicated support in the first decadal survey of the 21st century, and certainly a major factor in how smoothly it moved from proposal to ongoing project.

    Elsewhere, other observatories continued to move forwards, if in LIST’s large budgetary shadows. While Artemis missions had begun to explore the potential of the Moon as an astronomical platform, the establishment of the semi-permanent Orion moonbase had allowed Earth’s sister body to become a fledgling, yet in some ways important, astronomical center. Beginning with the establishment of the FROST-2 dishes shortly after the base’s construction, subsequent Orion missions continued to carry more specialized and larger-scale astronomical hardware, too expansive or complex to be suitable on a expedition mission yet too experimental or valuable to be deferred to a more permanent base. Among the first of these payloads were the first elements of the Lunar Low Frequency Observatory, a project seeking to take advantage of certain unique properties of the Moon to enable observations impossible from Earth. Although the ability of the Moon’s bulk to shield observatories on the Moon’s farside from the home planet’s intense radio traffic is well-known, less publicized is the fact that ionospheric heating and auroral waves prevent terrestrial radio observatories from receiving radio waves below about 20 MHz, leaving Earth’s astronomers completely blind to a vast swath of radio spectrum, and ignorant of what physical processes might be described by it. Only an observatory in space could detect the hum of these long wavelengths in the sky and help relieve that ignorance, but no space-based radio telescope capable of exploring these frequencies had ever been launched.

    The reason was quite simple. Although simple dipole antennas, much like those that had served the first radio astronomers, would be more than sufficient for detecting the massive electromagnetic waves being looked for, a large number of them, spread over a large surface area or volume of space would be necessary for conducting precision observations of the low-frequency sky, and in Earth orbit would still be disturbed by terrestrial low-frequency sources. With a complex implementation and a somewhat speculative payoff, the idea had never even come to the proposal stage. Only with the Artemis missions, and the possibility of a future Moon base, was it raised, as astronaut labor, or at least carefully teleoperated robots, could be used to emplace the thousands of antennas needed for a proper array, with the antennas themselves forming part of the payload of one or several cargo missions. The Artemis missions themselves were unsuitable for experimentation, as implanting and connecting the antennas would take up all or more than all of the available surface time, but the semi-permanent Orion moonbase was perfect for the project, with emplacement spread over weeks instead of days, and the concept quickly began development with the beginning of the Orion project. Soon after the base itself was established, containers of dipole antennas began making their way up to the base site, to be slowly removed and planted in the ground, one by one, gradually allowing astronomers on Earth to begin building a picture of the low-frequency sky. LLFO observations continue to this day, working in conjunction with high-frequency measurements like those conducted by the FROST-2 dishes and Earth-based low frequency instruments to provide a more complete picture of the sky.

    Alongside the Low Frequency Observatory, astronauts on the Moon were also deploying the next “advanced” lunar astronomy experiment. While the earlier Lunar Infrared Fixed Telescope had shown that an infrared telescope in the Moon’s permanently shadowed craters could function well, its design had limited its scientific value in favor of sheer mechanical simplicity, focusing completely on demonstrating the concept. The appearance of astronauts, at least periodically, able to repair and maintain potentially finicky low-temperature mechanical components meant that a more complex and scientifically useful telescope could be practically deployed with Orion than would have been possible for Artemis missions. From the very beginning, the scientists and engineers behind LIFT had hoped that a LITT—a Lunar Infrared Tracking Telescope—could be deployed at a future lunar base, and as soon as the Orion program began so did development of LITT. Taking the lessons learned from LIFT and merging them with a telescope of a more conventional optical and mechanical design, LITT was intended to serve as a perfectly conventional infrared telescope, conceptually similar to the array of instruments that had been built on Earth since the 1960s but with the twin advantages of a permanently cryogenically cool environment to improve sensitivity and the absence of infrared-absorbing atmospheric gases to impede observations. While considerably smaller and much less capable than the parallel LIST design, it would also be cheaper to build than that instrument, and could be operated for an indefinite period, periodically being upgraded with new instruments and technology. It would also have the advantage of having a virtually unimpeded view of the southern sky, something most Earth-based telescopes—located in the northern hemisphere—lack. Although LIST, too, would have such a view, and one of the northern sky that LITT would lack, time on the big instrument would be fiercely contested, and a smaller instrument would be just as good for many research programs. Since its emplacement during Orion 3, LITT has been a quietly valuable, though not revolutionary, scientific instrument.

    While astronauts on the Moon were emplacing LLFO and LITT, LIST was moving towards its own apotheosis of sorts. After over a decade of planning and design work, construction might have been expected to go smoothly, but like NASA’s previous great optical observatory, Hubble, LIST has had more than its fair share of trouble. Despite conservative budgetary estimates, unexpected problems in manufacturing the telescope’s main mirror, a beryllium alloy selected for its reflectivity in the infrared and stability against thermal stresses delayed construction and sent the telescope’s price tag upwards, grinding the project against the expense of constructing an operating the Orion moonbase and continuing to maintain Freedom. Other problems in the spacecraft’s bus, essential for providing power, communications, and pointing control; in diplomatic and technical negotiations with the European Space Agency and JAXA, participants in the telescope’s development; and in the complex deploying sunshade system, needed to shield the ultra-sensitive detectors from the heat of the Sun also contributed delays and unanticipated costs to the program. As the third-largest single line item in NASA’s budget, LIST’s overruns had an outsized effect on the rest of the agency’s programs, mostly by reducing (though not, fortunately, eliminating) the number of smaller, more specialized observatories that NASA launched during LIST’s construction.

    After all of that, its launch last year was fortunately and blessedly anticlimactic, with its Saturn-Centaur launch vehicle lofting it into space with all the quiet reliability and performance that launch vehicle family has built a reputation for. After a month of travel, the observatory reached its final observation point around the second Sun-Earth libration point, joining the aging but still functional Leavitt and several other observatories from Europe and Japan in one of the newest telescope clusters in the solar system. Although observations have only recently started, results from the observatory’s calibration and testing period indicate that it is fully functional, and, with a mirror nearly five meters in diameter, is now the largest space telescope in history, with an unparalleled sensitivity to infrared emissions. Lyman Spitzer, were he still alive, would surely be as pleased to see the telescope that now bears his name returning terabytes of scientific data to astronomers on Earth as he was to see the launch of Hubble and Leavitt, and as excited about the prospects for future researchers.

    Unfortunately, Spitzer, were he alive, would also have to contend with a community of space astronomers more divided than anything he had ever seen. With the growth of fields like exoplanet observations and the discovery of dark energy, more and more large projects have been developed, each promising attractive scientific breakthroughs, and each conflicting with the others over available funding and resources. Half a dozen ‘large’ projects—projected to cost over a billion dollars apiece—had significant degrees of support during the last decadal survey process, ranging from an ambitious proposal to build a gravitational wave detector in space to merely replacing the aging Leavitt observatory with a larger, more capable successor. Such division resulted not in productive competition between good ideas and the eventual selection of one truly outstanding concept, but instead to infighting within the decadal survey committee and, eventually, to none of the large missions being chosen for development at all. Instead, the decadal survey recommended that NASA continue and expand its successful Explorer program of small and medium-sized astronomy missions, and “explore cooperation” with the European and Japanese programs on missions seeking to observe exoplanets and the effects of cosmic inflation and dark matter. Although this has opened a window of opportunity for some mission concepts, continuing debate and the lack of European or Japanese missions to cooperate with has prevented any firm large mission plans from beginning. Instead, the American astronomical community remains in a state of uncertainty, unable to continue the program of large observatories that has made NASA the operator of some of the world’s most capable and desirable telescopes for the past thirty years.
     
    Part IV, Post 21: American unmanned exploration
  • Eyes Turned Skywards, Part IV: Post #21

    Even as their colleagues elsewhere were fighting to keep the Pioneer program going, the scientists and engineers working on the Hermes Mercury orbiter mission were steadily working towards their mission’s launch. Optimistic and aggressive early projections of launching in 2005 had run into the cold--or in this case, hot--reality of dealing with the harsh environment of Mercury orbit and forced a slip, but after recalibrating expectations mission development was going smoothly and on-time. The largest impact had been the addition of an Earth flyby before beginning the main flyby sequence to reach Mercury orbit, an optimization that permitted a slip of only seven months instead of nearly two years to the next Venus launch window. This would, admittedly, increase mission duration and risk, but the extra time would be spent in a relatively calm and quiet region of space, hopefully keeping the spacecraft safe and operational.

    Unfortunately, budgetary impacts were not so limited; as the first and possibly the only spacecraft to reach Mercury orbit for some time, the Hermes proposal had included a relatively large number of instruments in order to probe the entire geophysical environment of the planet. Although most of the instruments had been relatively well understood and developed, the complications of modifying them to deal with the Mercury environment and integrating them into the spacecraft had proven more difficult than expected, another reason for the mission’s slip. To avoid further delays, additional funds were poured into the project, one of the factors behind the delayed 2006 mission selection. Although effective at keeping the program on-track, the cost growth of the mission had led to questions during the reorganization of the Pioneer program about whether it was meeting its goals of enabling cheap, rapid exploration, despite the spacecraft’s successful launch only a few months after the cancelled selection.

    As NASA organized an abbreviated selection in 2007 to compensate for the cancellation of the 2006 selection, the board responsible was acutely aware that future Congressional support could very well be riding on whether the mission they picked was a successful exemplar of the program’s purpose or bloated and delayed by unexpected cost growth and technical issues. In a stroke of what could be called fortune, however, all of the proposals for the 2007 selection were repeats of proposals from the planned 2006 selection, as it was not possible for other teams to assemble their proposals fast enough to meet submission deadlines, and in turn were usually submissions of proposals that had previously gone through at least one Pioneer competition, since only larger, better-resourced teams could invest in suddenly updating their proposal documents to account for the new date. As a result, when the board met to consider its choices, it was selecting from a list of well-studied missions with extensive costing data, leaving very few unknown unknowns to anticipate. Even with this confidence, however, the selection of the Mars Polar Lander, or MPL, indicated that the board was breaking for the safe choice over more adventurous alternatives.

    Proposed in every competitive Pioneer selection from 1999 onwards, MPL was exactly and precisely what its name indicated, a fixed lander, like the Vikings, designed to land near one of the Martian poles and investigate the near-polar environment of Mars. Landing constraints on all previous Mars missions, whether Soviet or American, had prevented spacecraft from landing near either pole, a disappointment to scientists who wanted to explore the only polar ice caps in the Solar System aside from Earth’s. Besides being interesting subjects of exploration in their own right, their growth and shrinkage over the course of the Martian year has a measurable impact on the Martian atmosphere, with nearly a third of the atmosphere’s entire mass freezing out during winters and sublimating back into gas during the spring, piquing the interest of climatologists, while they are surrounded by and contain many complex and interesting geological structures, like layered terrain and underground water ice deposits. An offshoot of the pre-Pioneer Mars/Phobos ‘98 Ares Program mission proposal, which was ended by Gore’s election and reorganization of Project Constellation, the JPL-led spacecraft would land during the northern summer, spending several months exploring the weather and surface of a polar site before dying as temperatures plunged with the onset of winter.

    Although the mission had seen a great deal of study and refinement in the decade and a half it had taken to go from proposal to selection, and promised to stick to well-understood technology and well-understood risks, for this very reason its selection immediately saw criticism from some corners of the planetary science community who thought it showed a stodgy, conservative attitude completely at odds with the purpose of the Pioneer program. Probes like Barnard, still three years from its encounter with Comet Encke, or Hermes, about to flyby Earth on the way to Venus, had pushed the envelope, venturing to new places and daring to do new things that would be impossible with more expensive missions. A spacecraft that would recapitulate probes from thirty years earlier hardly had their air of romanticism and exploration, even if it was venturing to a new destination on Mars; even, perhaps especially, if it was a firmer budgetary and schedule choice than more adventurous missions. Nevertheless, most members of the community, aware of the criticism that had been leveled at Hermes, understood NASA’s desire to stick to the tried-and-true, and lined up behind the selection, leaving the critics as persistent gadflies, but nothing more for the moment. As work at Pasadena got underway, however, and attention elsewhere turned towards the next selection, their position would become increasingly important and influential.

    Thus, as the next set of Pioneer proposals began to roll in the next year, the attitude of the selection board had moved from sheer conservatism towards an attempt at blending careful conservatism and adventurousness in a single mission, much like Hermes had done. Although selecting missions unlikely to overrun their estimated budgets was important, so was a sense of adventure. Accommodating both of these desires was difficult, with many worthy missions at hand, but ultimately the board decided that the best balance for the 2009 selection was offered by the Venus Atmospheric System Explorer, or VASE, mission, an update of 1978’s Pioneer Venus Multiprobe mission to meet a new era of technology and a new set of scientific questions. By using a well-understood mission profile and operating inside the envelope, the board felt that it posed few financial or schedule risks, while at the same time it offered attractive prospects for novelty by traveling to one of the harshest planets in the solar system, and one that no American spacecraft had traveled to in fifteen years.

    VASE itself had grown out of a series of studies and mission concepts developed by the Venus science community since the demise of VOIR, which had developed a consensus that the most important questions about Venus concerned its atmosphere. Despite the Eos Franco-Soviet balloon mission of the early 1980s, the Pioneer Venus multiprobe mission a few years earlier, and even earlier Soviet atmospheric probes, much was still unknown about the structure and fine composition of the atmosphere, and therefore about the history of the planet. Compared to orbiters like the recent Japanese Akatsuki and Chinese Jinxing spacecraft, atmospheric probes could explore the denser, higher-pressure region under the cloud deck, providing otherwise unobtainable information and complementing orbital observations. Similarly, a modern balloon mission, using more advanced electronics and equipment than the old Eos mission, could provide more information on the atmosphere in more places at the same time. Despite this scientific attraction, atmospheric missions were also much less ambitious than many of the alternatives, avoiding the harsher regions of Venus or the need to survive them for long periods of time.

    However, even lower ambition only went so far in reducing cost and schedule risks. Although the Venus Exploring Inflatable Lifter, or VEIL, mission had proposed a Venus balloon and at first glance seemed a strong competitor to VASE, it had been heavily compromised in order to fit in the Pioneer budget box. Venus scientists wanted multiple balloons at multiple latitudes and longitudes in order to track wind currents at different locations on Venus, grist for the mill of computational models that they could then use to extend their observations to the rest of the atmosphere. In addition, they wanted the ability to track balloon movement anywhere on the planet, in turn requiring a communications relay orbiter for when the balloons were inconveniently on the side facing away from Earth. Together, this meant much higher costs, in the range of a Cornerstone mission, so that VEIL had pared the mission back dramatically, only delivering a single balloon with no communications relay satellite. Although this managed to reduce the mission’s projected cost to acceptable levels, it also reduced its scientific value so much that it hardly seemed worthwhile any more, while still throwing up a number of red flags for potential overruns and schedule slips in the making. As much as many people wanted the Pioneer missions to push the envelope, they could only push so far. Ultimately, even one balloon had proved a bridge too far for Pioneer’s budgets, and VEIL had lost out in favor of VASE.

    For more ambitious missions, scientists would have to look towards the Planetary Science Prioritization Panel, or PSP^2, as it began working on its legally-mandated report on the future of planetary science soon after the selection of VASE. Since the 1960s, the National Academy of Sciences COMmittee on Planetary and Lunar EXploration, or COMPLEX, had periodically produced reports on recommended future planetary science missions, many of which had eventually been flown, but in a haphazard and demand-driven process. Unlike other expensive fields like particle physics, nuclear physics, or astronomy, there was no statutory structure for providing recommendations to guide and shape NASA’s planetary science program, with the result that it had evolved to launch missions based at least as much on political as scientific criteria. This had worked well in the afterglow of Apollo and the Vulkan Panic, when relatively loose budgets had permitted a wide-ranging and pioneering program of exploration, but had become ineffective in the more straitened circumstances of the 1990s. One of the primary recommendations of the Cornell Plan had been to establish a permanent advisory body like the Particle Physics Project Prioritization Panel or Nuclear Science Advisory Committees advising the Department of Energy and National Science Foundation on their fields, but for planetary science. This body would be responsible for regularly providing a list of recommended missions to NASA, which the agency could then choose from. With its establishment by Congress, PSP^2 was now beginning to draw up its first report with the development of studies on a wide range of possible missions and destinations by specialized panels including many of the most distinguished scientists in the field. Each of the five major groups of target bodies--terrestrial planets, Mars, gas giants, icy bodies, and minor planets--had its own survey panel, along with “interdisciplinary” survey panels covering technology development and supporting infrastructure. Over the course of the year, PSP^2 sorted out the responsibilities of the panels and began a concerted program of community outreach, even as the panels, in turn, began to write up their own studies for the main committee.

    As the first drafts of the panel studies began to come in later in the year, it became clear that PSP^2 would, at least, not suffer from a lack of mission possibilities to choose from. Over two dozen serious proposals had been mooted and studied by the target panels, ranging from a large mission to follow up Hermes and perhaps even attempt a landing on Mercury’s surface to an ion-propelled, Saturn Heavy-boosted spacecraft that would zoom past Jupiter before screaming out into interstellar space at a clip far exceeding either the Voyagers or the earlier Pioneer probes. Along with the variety of missions, cost estimates fluctuated wildly: the proposed Mars Geophysical Network came in only a little above the Pioneer cost cap, while others, like the behemoth Neptune-Triton System Mission, promised to break Cassini’s three billion dollar record and become the most expensive planetary mission in NASA’s history. With such a wide range of proposals, the most pressing question had become how to evaluate and rank so many missions.

    Even as PSP^2 was considering how to manage the flood of proposals, MPL’s launch window was finally rolling around in October 2009, over half a year since VASE’s approval. Fitting for a mission that had fifteen years of design heritage behind it, it had experienced few difficulties in development, with comparatively smooth sailing on all fronts whether schedule or budgetary. Launch was equally smooth, though marked by a major first for the space program; instead of the traditional Deltas or Saturns, MPL was lifted into orbit by a Star Launch Thunderbolt, with a solid third stage attached to the spacecraft’s aeroshell completing the injection to a fast Mars-crossing trajectory that would reach the Red Planet in just six months instead of the usual eleven, so that MPL would reach the Martian north pole just as spring was turning into summer. The successful launch marked the first time that a NASA planetary science mission had ever been launched on a reusable vehicle, and a significant cost savings compared to more conventional vehicles. After launch, MPL settled into dormancy as it cruised towards Mars.

    While one of the most recent Pioneer missions was nearing Mars, one of the oldest was finally reaching its destination after an epic, decade-long journey. Although reaching Encke had been time-consuming, the reward was sweet, as after multiple flybys of Venus and Earth to pump down the spacecraft’s orbital energy, Barnard became the second spacecraft in history to land on a comet in March of 2010, a month after reaching and putting itself into orbit around Encke. After a tense landing sequence, punctuated by a brief scare when it appeared that the mechanism designed to anchor the spacecraft to the comet had malfunctioned and Barnard had bounced back into orbit, it quickly set to work. Although the first images sent back from the lander showed jagged spires of relatively solid rock and other materials jutting ominously out of the comet’s surface, leading commentators to compare Encke to Mordor from J.R.R. Tolkien's Lord of the Rings books, further exploration over the next few weeks showed that this outwards appearance was only a surface illusion. Instead, most of the surface was covered in a thick, refractory layer of rocky dust, extending tens of centimeters into the comet and protecting its icier core from the harsh conditions of space, making the comet a kind of “icy dirtball,” as Helios-Encke some thirty years earlier had indicated. Barnard also built on Kirchhoff’s data on Tempel 2, showing that Encke-type comets also had little if any intrinsic magnetic field and an elevated level of “heavy hydrogen,” deuterium, compared to ordinary hydrogen, providing more evidence against the theory that cometary impacts had supplied most of the Earth’s water.

    Just a month after Barnard touched down on Encke, MPL entered the atmosphere of Mars, streaking through nearly cloudless skies before safely descending to a landing point near the planet’s north pole. In a spectacular example of inter-agency cooperation, Japan’s Hayabusa[/] orbiter, which had only recently reached the planet itself, managed to successfully image the probe while it was descending under its parachute, the first time a spacecraft around another planet had recorded another, let alone while it was passing through the atmosphere. After this high note, MPL’s mission continued from success to success, with the probe’s robotic arm quickly revealing hoped-for evidence of water ice just under the Martian surface and the chemistry lab definitively confirming that perchlorates, believed to be responsible for Viking’s inconclusive biological results, were present in Martian soil. Although a blow to those few who hoped life might be present there, most scientists were pleased that this mystery had finally been definitively resolved, and focused more on the geological and climatological implications of MPL’s other data. MPL’s meteorological suite also provided important information on ground-level conditions near the Martian poles, complementing Viking and Mars Traverse Rover point data and orbital observations of the entire globe to help build refined atmospheric models, both of Mars and of Earth. After nearly six months of observations, MPL finally succumbed to dropping temperatures and dimming sunlight as Mars neared its autumnal equinox, shutting down for the final time in late September of 2010. Hayabusa observations of the site showed that dry ice formed around and over the probe that winter, as expected, and most likely destroyed MPL’s solar panels from having to support excessive, beyond-design weights. Between that and cold damage to the electronics, the spacecraft was permanently disabled, explaining why attempts to revive the spacecraft during the next Martian summer in summer of 2012 were unsuccessful.

    Meanwhile, PSP^2 was nearing the completion of its final report after a long sequence of community outreach events, communications with NASA and other space agencies, and internal studies. Based on the scientific priorities it had drawn up with involvement from the wider planetary science community, the panel was generating a list of broad recommendations supplemented by a ranking of mission proposals according to cost and scientific value. Finally, after nearly two years of work, it published its draft final report in early 2011, prompting a final round of feedback before the publication of the final report that summer. As might have been expected after the Cornell Plan, it urged NASA in the strongest tones to launch at least one larger mission over the next decade and continue the Pioneer program with a cadence of one mission selection every other year, along with establishing an intermediate-cost program for missions like the Mars Geophysical Network or Venus Atmosphere Circulation Mission that were too large for Pioneer but, at least in the judgement of the committee, too small for a full-scale Cornerstone mission. Additionally, it recommended that NASA establish an ongoing technology development program rather than leaving it up to each individual mission to pay for whatever new instruments or technologies that it might need, and for NASA to seek more international partnerships with other space agencies. Finally, for the next Cornerstone mission the committee recommended that NASA immediately begin work on a Europa Orbiter mission, especially if further study could reduce the estimated price tag, followed by a Mars Sample Return campaign later in the decade.

    Crucially for this recommendation, icy moons and Mars scientists had held the balance of power on PSP^2 and in the wider community; the United States had launched more missions to Mars than to any other planet or heavenly body aside from the Moon, and Galileo and Cassini’s missions to Jupiter and Saturn, respectively, had maintained a healthy community of scientists interested in Europa, Titan, Enceladus, Triton, and other, similar bodies. Compared to the relatively small and marginalized Venus or ice giants communities, icy moons and Mars had a large number of advocates able to push their case and develop mission concepts, while being simpler in a number of respects than missions to these other locations. Additionally, Mars and icy moons had built-in public relations advantages; Mars had been the destination for interplanetary spaceflight of any kind for over a century, while the icy moons, especially Europa, were widely regarded as the most likely place for alien life to exist in the solar system, both tempting attractions for the public and for planetary scientists. Although advocates of other destinations could make powerful arguments in their favor, they lacked the sheer magnetic draw of Mars or Europa as destinations, and had fallen behind in advocacy despite their best efforts. The result was a compromise between these two groups, based on the fact that the Europa Orbiter would be a cheaper and simpler mission than the multi-mission Mars Sample Return campaign JPL had envisioned since the 1980s. Combined with their recommendation of an active Pioneer program, which could easily support smaller Mars missions, Venus missions, and minor planets missions, and an intermediate category that could cover larger and gas giants missions, the final report had something for just about everyone in the planetary science community, thus preserving the unified voice that had led to the creation of PSP^2 in the first place.

    While Congress had already approved and funded an ongoing frequent Pioneer program, the board’s other recommendations would require approval from the Hill, so that the publication of the report in early 2011 only marked the beginning of a new advocacy campaign. As they had in previous years, planetary scientists alternated pilgrimages to Washington with work in the lab focusing on addressing Congressional concerns with their recommendations. Despite the estimated cost of a Europa mission, Congress proved receptive to the idea of a new mission focused on Europa’s postulated under-ice sea, perhaps inspired by the idea of discovering alien life, though they had concerns over the cost. Despite being operationally simpler and having a projected cost less than Mars Sample Return, the Europa Orbiter was still expected to cost considerably more than Cassini, a result of the complex mission requirements, the difficulty of reaching Europa orbit, and the extremely harsh conditions of near-Jovian space, especially the radiation environment. Prompted by these Congressional concerns, interested scientists refined the mission concept, dredging up an idea from the early days of Europa mission planning in the 1990s to have a Jovian orbiter make many Europa flybys instead of actually putting itself into Europa orbit, thus reducing the amount of propellant needed and simplifying the mission. Moreover, by spending only brief periods of time deep inside the Jovian radiation belts instead of being forced to continually reside in them, the lifetime of the spacecraft could be greatly increased and it could perform some incidental science focused on the outer moons Ganymede and Callisto, and on Jupiter itself. NASA study of the concept also showed that this option’s impact on science would be minimal or, in some cases, actually positive compared with the orbiter option, while the cost of the mission could be slashed by more than half. With such positive results, Congress approved a new start on the Europa Systems Mission, or ESM, last year, with a Saturn Heavy launch planned to boost the probe towards Jupiter in 2019 or 2020.

    In the meantime, however, other new recommendations by PSP^2 suffered. Despite the panel’s strong insistence on a new intermediate-class program to fill the gap between cheap Pioneer missions and expensive Cornerstones, Congress proved cool to the idea and it was dropped by advocates shortly after the report’s publication, as were the establishment of a stable technology development budget line and the creation of a budget line for supporting data archiving and release and facilitating the development of new planetary scientists. Even the projected Mars Sample Return mission suffered, with serious study work not beginning until after the approval of ESM in 2014, and no flight now expected until the mid-2020s at best, possibly slipping mission completion to the late 2020s or even the early 2030s at earliest, about when some of the more optimistic recent projections put humans on or around Mars.

    On the other hand, the Pioneer program has continued from success to success; even as PSP^2’s final report was being published, the Pioneer selection board announced that it had chosen the Mars Ice Orbiter, or MIO, for launch in 2013. Building on the detection of ice at MPL’s landing site, MIO is designed to use a ground-penetrating radar system to detect ice all over the planet, expanding MPL’s observations towards a global inventory of Martian subsurface ice. Since its arrival last year, MIO has been aerobraking into its final observation orbit, which it is expected to reach in October. In the 2013 selection, the selection board opted to return to the minor planets by choosing the Comet Tour mission, a multi-flyby spacecraft utilizing electric propulsion to visit several cometary nuclei to compare them with Encke and Tempel 2 and build a broader baseline of cometary properties for comparative paleontological analysis. Data from 2010’s private NEOSearch spacecraft has been very useful for constructing a target list, and launch is scheduled for later this year. Most recently, this year’s selection broke the pattern of alternation between Mars and other targets that had been beginning to draw some criticism from planetary scientists, instead opting for the Aeneas mission to the Trojan asteroids. Clustered before and behind Jupiter in its orbital path, the Trojans are believed to be relatively volatile-rich remnants of the early solar system, and windows into our solar system’s past, so that Aeneas may shed important light on the formation of the gas giants and other outer system bodies.

    Farther afield, planetary scientists are already gearing up for the next PSP^2 survey, expected to begin in 2019, and for the Mars Sample Return mission planned for the next decade. Recent MSR studies have shown a need for at least one precursor mission, given the large, sophisticated rover planned for use in gathering samples and the complexity of precision landing on Mars, so that support is growing for a large rover mission in the early 2020s, following the launch of ESM, while cost issues have led NASA to approach ESA, Roscosmos, and JAXA about cooperating on the sample return itself. The idea of creating an intermediate mission class has also returned, though as of yet it lacks significant Congressional support. Nevertheless, should their new process continue functioning, the future of planetary science looks brighter than it has in decades, with a steady stream of missions set to head skywards and explore new--or old--worlds.
     
    Part IV, Post 22: NASA after Orion and the Saturn-II
  • Eyes Turned Skywards, Part IV: Post #22

    Although 2008's Orion 1 mission appeared to be just another in the series of American-led lunar landings that had spanned the last decade, under the surface it represented a significant advancement from the Artemis missions. The introduction of the new pressurized rover enabled the crew to complete several multi-day traverses, including a 66 kilometer circumnavigation of Shackleton crater which hosted Lunar Outpost Orion (informally known, among other names, as “Shackleton Base” or “the Shack”). Additionally, the flight saw the successful debut of the Russian Luna-Pe lunar logistics vehicle, proving that (at least for the moment) they had excised some of the demons haunting their unmanned program and demonstrating a path forward for more ambitious future missions. After the return of Orion 1's crew, mission controllers continued to monitor the outpost's system as training of Orion 2's astronauts stepped up, preparing for the day that they would make the first return to a lunar base in history. This second mission would build on the infrastructure already established at the Shackleton site, spending a full three months on the lunar surface at what was, inevitably, described by the press as the moon base. In the course of this extended stay, the crew were to expand the site survey already begun by Whitt’s Orion 1 crew, venturing on multi-day traverses to survey other polar craters’ geography and test for the presence of frozen volatiles such as those already identified within Shackleton. However, they would also make extensive use of the outpost’s biology and medical laboratory capability, using the mission’s extended duration (nearly as long as all of the original Artemis missions combined) to test the effects of longer stays in lunar gravity on the crew, for comparison for baseline data already collected on Spacelab and Freedom, and rat studies in the pseudo-gravity of Freedom’s centrifuge lab.

    The experiment which attracted the most attention during the mission, though, was surely the “moon farm,” aimed at evaluating the growth of plants brought from Earth in a variety of lunar conditions, including Earth-sourced soil, treated lunar regolith, and a primitive hydroponic setup. After the crew’s arrival at Shackleton in March 2009, NASA public affairs encouraged educators to have their students “follow along” in classrooms back on Earth, comparing the growth of their own beans, herbs, and potatoes to those sprouting on the moon. While Edward Boxall had been the primary “outreach source” for Orion 1, Orion 2’s main face in the media was thus Mission Science Officer Hannah Parker. With proper treatment and enrichment with bacteria and biological media, the lunar regolith proved to be a viable growth medium for crops, and the crew’s pictures of digging early-sprouting potatoes and carrots from the beds would be seen almost as often as images of their record-breaking 5-day trip to Sverdrup and de Gerlache craters. Sadly, though, the astronauts wouldn’t get to taste either the fruits or vegetables of their labor, as the harvest was to be returned to Earth for nutritional and chemical analysis to ensure the suitability of future lunar crops for human consumption.

    While the first two Orion missions were proving tremendously successful in terms of scientific return--particularly given that each actually cost less than an Artemis sortie mission--plans for the future of NASA exploration were more up for debate. Orion, after all, was intended as a short-term outpost, incapable of effectively supporting small crews for more than a few months at a time, nothing like the more extensive (and more self-sufficient) lunar outposts being studied by Oasis project teams. At the same time, Space Station Freedom was beginning to reach the end of its design life--the station had already had its lifespan extended to 2014 by planned replacement of batteries and onboard systems, but eventually NASA would need to replace both outposts. Additionally, the time since either Orion and (certainly) Freedom had been planned had seen major advances in the state of the art in spaceflight, and further changes seemed on the horizon. The completely expendable Saturn Multibody seemed outdated when compared with the flight-proven semi-reusable Thunderbolt, much less beside the Lockheed Starclipper and European Aquila two-stage fully reusable rockets under active development. The potential of capturing some of these commercial and technological benefits in any full-scale lunar outpost or replacement for Freedom was immense, and many within NASA viewed a new wide-ranging evaluation of the agency’s programs, vehicles, and goals to be necessary to maintain the agency’s leadership into the coming era of spaceflight. In 2009, as the Orion 2 crew worked on the ground at Shackleton, the newly re-elected President James Woods heeded these thoughts, directing Administrator Banks to conduct a broad-scope review of NASA’s capabilities and direction, and provide recommendations for a plan for the future, similar to the Richards-Davis Report from more than 15 year earlier.

    The report began its evaluations with the fundamentals of the agency’s human spaceflight capabilities: Apollo and Saturn. For almost 50 years, pairings of these two families had served as NASA’s backbone, from the original Apollo program to Skylab, Spacelab to Freedom, and Artemis to Orion. However, Saturn Multibody--for all its flexibility and relatively low costs for an expendable rocket--was more than twenty years old, and even the “Interim Improvement Program” variant which lifted crews to Freedom and Orion dated to the Gore administration. Placed beside Thunderbolt’s reusability and low cost, the Saturn seemed exorbitantly expensive, and even since 2004 NASA had been considering a reusable replacement heavy lifter. The initial “Phase A” round of studies--small and cheap conceptual evaluations from a wide variety of contractors--had returned in 2006, and ranged from “simple” additions of reusable engine pods or flyback hardware to the existing Saturn core, to immense clean-sheet reusable first and second stages capable, like Saturn, of throwing more than 70 tons into Earth orbit. However, while the agency generally considered some sort of new, reusable launch vehicle critical to enabling a revamped exploration program continuing in LEO, on the Moon, or even beyond, the Phase A proposals ran into a profound case of “sticker shock.” Mindful of his boss’ budget-conscious ways, Banks had commenced Phase B studies in 2007, focusing more on cost-effective ways of implementing at least some level of reuse while retaining Saturn-class performance, striving for a compromise that would displease as few as possible, even if it wouldn’t perfectly satisfy anyone. Thus, when the Banks Report landed on the President’s desk in late 2009 for evaluation of options and development of a cohesive plan with Congress on the Agency’s future, it already came with a suggestion for the new backbone vehicle for the agency’s future: the Saturn II.

    As its name suggested, the Saturn II was a continuation of the agency’s Saturn line, using the same 6.6m main tankage and clustered-core design as Saturn Multibody to reach the required levels of payload to launch new space stations and support the heavy and bulky payloads under study for Oasis. However, the vehicle would also change many details in pursuit of more cost-effective operations, starting literally from the ground up. When designers at Boeing had considered the addition of reusability to the Saturn Multibody family, they were faced by the problem of trading added weight--in both structure and reusability systems--against the system’s payload capacity--its reason to be. However, since reusability of at least the first stage was viewed as critical to ensuring Saturn II’s cost effectiveness, Boeing turned to new technologies in pursuit of a solution. Staged combustion kerosene engines had been something of a novelty in the United States since the technology had made its way out of Russia after the collapse of the Soviet Union, and the more-efficient, higher-performing engines had seen several focused development programs during the Gore and Richards administrations. To gain reusability without sacrificing capability, Boeing proposed to replace the venerable F-1 family of Saturn V, Saturn IC, and Saturn Multibody with a pair of staged combustion engines, the Rocketdyne RS-76, a US-developed engine incorporating the latest in American production methods and the secrets of Soviet engine technology. Two of these engines, each producing 4 MN of sea-level thrust, would be the basis of the new Saturn II core. The margin bought by their much-improved specific impulse provided the cushion for adding reuse to the vehicle.

    This reuse system had been a point of debate, even within the Boeing design team, with some favoring addition of wings and jet turbines, as with the European Aquila system which became public in the same year Saturn II was proposed to NASA. However, the difficulty was that a tremendous increase in structural mass would be necessary to accommodate the loads that could be expected in two axes for vertical takeoff and horizontal powered flight and landing, particularly given the tightly-optimized low weight of the basic Saturn Multibody tanks. With the more efficient main engines, Boeing had proposed--and NASA had approved--an initial Saturn II configuration which would return to its launch site and land in the manner of the StarLaunch Thunderbolt, indeed even proposing to partner with Allen’s SLS on the development of the new core. However, one problem complicated this: the thrust of the new main engines. While this would enhance their ability to perform the initial “boost back” burn and slow the stage’s entry into the atmosphere, there was no chance of achieving sufficient throttle authority for a safe touchdown on these engines. Therefore, for final touchdown, Boeing proposed to use a simple pair of smaller gas-generator engines, filling gaps on the vehicle boat-tail. These engines would be easier to re-light mid-flight, and their lower thrust would make the vehicle not only capable of landing safely, but also of achieving a T/W of 1 during touchdown under minimum throttle, enabling a gentler touchdown than Thunderbolt. The weight added by these landing engines was to be partially compensated for by using them during ascent for all the vehicle’s pitch, roll, and yaw control, meaning that the main engines could be fixed in place--a savings in gimbal mass and complexity judged by Boeing’s engineers to be more effective than a massive clustering of smaller staged combustion main engines.

    In operation, the Saturn II proposal would launch from the ground in either a single or triple-core configuration (support for solid rocket boosters was to be dropped, with the gap filled by the cheaper reusable liquid boosters). The second stage was to be either the S-IVB or the S-IVC of the Multibody with minimal updates, depending on the mission, given that the costs of developing a large reusable orbital stage had been judged prohibitive within NASA’s cramped budget. However, to further minimize the penalties to their new heavy’s capability, NASA would also invest in new infrastructure--a barge or sea platform capable of serving as a downrange landing site for high-capacity single core flights or for the center core of Saturn II Heavies. The result of this and the more efficient main propulsion was that in spite of the penalties of reuse, Saturn II would have similar capabilities of the vehicle it was replacing: 21 tons to low Earth orbit on a single core and 76 tons in a Heavy configuration with downrange recovery of the center core, but with almost half the cost per launch. For those who had hoped for a fully-reusable super-heavy, Saturn II looked like a half-hearted compromise next to graphics of Aquila, Starclipper, or proposed Thunderbolt reusable second stages. However, NASA’s budget could only stretch to cover so much new development, given the ongoing lunar and LEO outpost program, and Saturn II was an implementation of relatively-conservative technology well-proven by Thunderbolt L1 that would give enough of the benefits of reuse to answer NASA’s needs.

    Those needs certainly did not end with the new launch vehicle--instead, Saturn II was to be just a part of using the existing human spaceflight budget more efficiently to cover a more expansive program. Just as Orion’s outpost design enabled exchanging 3-launch $1.5 billion annual sorties lasting two weeks with $1.2 billion 2-launch missions lasting several times as long, Bank’s suggestion was to exploit reusable launch vehicles to do even more with the budget of Freedom and Orion. In the immediate future, he envisioned three major tracks, addressing three exploration missions--a deliberate nod to the “Constellations of Exploration” of Woods’ most immediate Republican predecessor. In Earth orbit, the Bank Report called for a new space station to replace the aging Space Station Freedom’s scientific capabilities. This could consist of one or two large modules launched aboard Saturn II, creating a basic station core which could then be supplemented in an evolutionary fashion by rigid or expandable modules launched to provide increased power, laboratory, and habitat capacity as-justified by the intentions of NASA, international partner, or commercial firms. In addition, he proposed that NASA’s venerable Apollo be merely the contingency capacity for this station, with primary logistics (if possible even including crew) to be contracted out to US commercial operators or bartered to international partners. This would leave the station as little as half as costly to operate as Freedom, a savings which could be rolled into the Oasis expanded lunar outpost.

    Here, too, Banks called for supplementing Saturn II-launched primary systems with commercial capacity, and drew on plans previously studied by NASA and--particularly--Northrop. The overall plan for a sustainable lunar outpost was for Saturn-launched payloads and descent stages to convey crews and large components to the surface by drawing on a commercial-fed network of Pegasus-derived, TransOrbital-style tugs. With a single Saturn II-H launch, an Apollo and descent stage could be placed into orbit, as with the crew launch in Artemis in Orion. Here, though, the Pegasus tug to boost them to EML-2 would already be waiting in Earth orbit, topped from NASA depots kept filled by a competitive propellant market. At EML-2, a smaller depot (a derivative of Northrop’s existing Centaur-based TransOrbital depot) would allow the Pegasus to partially refill and begin the transit to lunar orbit, acting as a partial “uncrasher” before leaving the crew to their mission to the surface and returning via the EML-2 depot to Earth for reuse. The end result would be dramatically improved payload to the lunar surface while at the same time reducing cost. This, in turn, would be used to stretch each crew rotation to six months, and launch two missions a year--transitioning the Orion outpost site to a permanent Oasis facility for the same annual budget. Because these two prongs of the Banks plan could be implemented by redirection and more efficient spending of NASA’s existing budget, achieving a lunar base and new space station could serve as a foundation for more if the agency’s budget saw expansion in the future.

    Banks knew his boss well, and the notion of leaning on privatized routine launches to reduce public spending while enhancing capability was almost a rhyme with Woods’ more general philosophy about government. Though a tough sell in Congress, ultimately the plan found enough support to pass; those representing NASA-heavy districts saw none of the feared cuts to their constituents’ livelihoods, while budget hawks like the President could hardly complain--even if the plan didn’t cut NASA spending, it at least wouldn’t require increases. The plan also proved popular with the public, where the new station and “first permanent lunar base” briefly captured public attention. Contracts for the development of the new station core, the Pegasus tug and depot hardware, and the new Saturn II were issued, supervised by the staff at Marshall, while Johnson settled into defining programs for the commercial supply of supplies, propellant, and potentially even crew to the next-generation exploration system.

    While NASA was in the reviewing its goals and determining the best mix of private and government launches to enable the nation’s spaceflight goals, other commercial applications were being explored based on the same reduced cost of access to space. In addition to these long-speculated applications, new, more novel avenues were being explored than second-generation LEO communications constellations and the wide variety of plans for space tourism that had been fermenting since Thunderbolt was introduced. Funded in part by private investors and in part by the NSO, NEOSearch (founded by, among others, Peggy Barnes, ex-NASA astronaut) has been among the most notable and most successful of these efforts, launching one spacecraft five years ago and currently planning a second.

    It was intended as a partially-charity, partially-commercial endeavor to launch a small, half-meter infrared telescope into Earth orbit, dedicated entirely to the mission of scanning the sky for asteroids. NEOSearch estimated that a single telescope in low Earth orbit would be able to catalogue over 80% of so-called “potentially hazardous objects,” or PHOs, over a five-year mission duration, blowing past earlier ground-based efforts to detect threatening asteroids, while costing little more than similarly efficient ground-based systems. However, in the process they would also amass the most complete database of asteroid properties, including orbital parameters, sizes, and compositions, in history--exactly the data which would be required to tap these “flying mountains” for rare metals and volatiles in the asteroid mining which had long featured in science fiction. Just as operators like Thunderbolt and Starclipper were turning science fiction into science fact, NEOSearch planned to sell access to their database to several firms popping up which planned to begin such asteroid mining in order to pay back the investments--a combination of global protection and capitalist success. After several years of fundraising, NEOSearch was able to begin bending metal, culminating in a launch on a Thunderbolt in early 2010. Over the past five years, the NEOSearch One telescope has become by far the largest single contributor of new objects to the minor planet lists, although the latest public data set is still estimated to be several percentage points short of the 80% PHO target.

    Nevertheless, the telescope is healthy and data collection and processing are ongoing, so it may make up that distance yet. In the meantime, analysis of the public data set has not revealed any actually hazardous objects, and although several startups have ponied up the cash needed to obtain detailed information, all of them seem some ways from launch. NEOSearch itself is studying how to improve detection of objects interior to Earth’s orbit, which comprise most of the undetected fraction of PHOs at this point, and which are difficult for a telescope on or near Earth to detect due to their proximity to the Sun. Currently published concepts include putting a telescope at the first Earth-Sun Lagrange point, ESL-1, which will marginally improve detection ability, and putting one in a substantially lower heliocentric orbit, similar to that of Venus, where detection ability would be greatly increased. Launch costs to that orbit would, however, be considerably higher, as would spacecraft temperatures, while communications would be slower and more expensive to achieve than in low orbit or even at ESL-1. So far, no choice between the two has been publicly announced.

    While pioneers of science fiction and spaceflight from Clarke to Von Braun might have been amazed by such a privately-funded science and exploration mission--separate from any national space programs and with a business plan half-charity and half-market-speculation--the explosion in even smaller unaffiliated payloads was astounding even to those witnessing it. In the early years of spaceflight, superpowers had mustered massive efforts of miniaturization and launch vehicle development to launch payloads like Vanguard, Explorer, or Sputnik, weighing less than 100 kg and consisting of little more than tight packages of batteries, solar panels, and radio antennas. Since then, most payloads had become much larger, but at the same time, the computer age had seen the minimum size of such “simple” systems as the original space vehicles shrink, and occasional proposals had circulated for the creation of “microsats”--satellites with minimum levels of capability but which could be extremely low-mass and, as a result, be developed and launched extraordinarily cheaply. Around the turn of the millennium, this concept had once again been current, and a joint MIT-Stanford initiative had yielded the “SpriteSat” standard--a tiny cubic satellite with rigidly standardized exterior dimensions, mass properties, and interfaces. Using little more processing power than was available in the burgeoning PDA or “smartphone” market, these spritesats could be assembled cheaply by universities, clubs, or even private individuals, then launched on any flight which could accommodate several kilograms of ballast. The concept had a slow start, but had rapidly exploded after 2005, fed by increasingly cheap electronics, availability of modular kit solutions, and the dropping cost of launch associated with NASA initiatives and the StarLaunch Thunderbolt.

    As spritesats proliferated, solutions to the form factor’s power, avionics, communications requirements became available “off-the-shelf” (in some cases incorporating modified versions of handheld “satphone” systems for basic communications and telemetry, or PDA or smartphone processors as onboard avionics), and some NASA and university groups began to experiment with the potential which could be crammed into such a small, modular platform, attempting to miniaturize scientific instruments, communications gear, and even propulsion to turn spritesats from simple tumbling boxes on slowly decaying LEO tractories into fully-functional science platforms, potentially capable of even leaving LEO behind. The result was a growing divide among users of the modular hardware between groups or individuals building simple “dumb” spritesats to simply have, for a brief time, the achievement of their own personal microsatellite and those pushing the bounds of the capabilities of the systems, some of whom began to talk about a new standard with larger mass and dimensional limits and correspondingly larger potential as scientific platforms that would still be able to exploit some of the advances already being made in spritesats.

    Most emblematic of this second group was the 2010 release of two ion-equipped NASA spritesats from the Orion 3 expedition’s discarded Pegasus stage after the completion of the crew’s Earth departure burn and the stage’s separation from the stack. Using limited onboard navigation, the spritesats--arguably tiny probes--were maneuvered from their near-heliocentric initial orbits, letting their maneuvering and ground-based signal tracking probe weak stability bounds in the outermost reaches of cislunar space. If advocates of larger microsats managed to make progress, the limited “Skipper” 1 and 2 (named for their track across the ripple-like shapes of cislunar gravitational potential wells near L-1, L-2, L-4, and L-5) might be followed in the future by more ambitious cheap missions.

    However, such ambitions for cost-effective missions weren’t just limited to small spacecraft. Building on their past legacy and their accomplishments with the private sector, NASA was able to launch renewed programs in both LEO and on the lunar surface, while meeting the President’s goals of only minimal changes in the agency’s budget. At the same time, the private sector was beginning to take its own steps in growing markets for Earth-focused space applications and beginning to look at space-focused programs. The potential of commercial and government cooperation was being proven again and again, on missions ranging from individual kilograms to hundreds of tons and with budgets spanning six orders of magnitude.
     
    Part IV, Post 23: Japanese space probes
  • Eyes Turned Skywards, Part IV: Post #23

    When Fukuro launched in late 2001, it was the end of a long era of Japanese planetary exploration. Since the launch of their first interplanetary probes fifteen years earlier, Japan’s planetary and comet exploration spacecraft had all been built to the same fundamental design drum-shaped design, based on communications satellite designs popular in the 1980s. Although variations on this basic design had served Japan well for fourteen years and four spacecraft, the design had been pushed to its limits for Fukuro, and it was obvious that new approaches would be needed for future missions such as the new Venus probe that Japanese Institute of Space and Astronautical Science, or ISAS, had recently begun.

    In a fortuitous but entirely planned convergence, at the same time it was beginning to think about Fukuro’s successor ISAS was also putting the finishing touches on its new launch vehicle, the Mu-5 or M-V, a replacement for the series of Mu-3 vehicles it had relied on since the 1970s. While NASDA had led the H-I and H-II projects to develop a new, relatively large all-Japanese launch vehicle, ISAS had been heavily involved in the development of their solid rocket boosters, comparable to those used by American Delta rockets due to its lengthy experience in building and operating solid rocket motors. In turn, from the beginning of the H-I program it had been planned that these solid rocket boosters could be redeveloped into a standalone booster with a greater payload capacity and lower cost than the Mu-3 series. Such a booster would provide an independent, cheaper alternative to the H-I for scientific probes and other small spacecraft, and would provide useful experience with large solid rocket boosters and motors that could, perhaps, later be adapted to military roles. With the entry of the H-I into service, ISAS had turned its full attention to completing the Mu-5, and by the time Fukuro was sent into space was on the verge of performing its first test launch.

    As Mu-5 development had started, ISAS had begun planning new missions that would take full advantage of the new booster’s capabilities, including Earth-orbital and deep-space missions. Besides a range of application missions focused on studying the Earth, Sun, and near-Earth environment, ISAS studied a broad set of beyond Earth-orbit missions, including lunar orbiters and landers, Venus and Mars spacecraft of several types, comet and asteroid flybys, landers, penetrators, and rendezvous missions, and several plans envisioning visits to more distant destinations, including Mercury, the asteroid belt, and even Jupiter. Preliminary study showed that only the nearer destinations were practical given available resources, both budgetary and technological, so that ISAS switched to a focus on what it termed its “three areas of interest,” the Moon, Mars and Venus, and the near-Earth comets and asteroids.

    By the time ISAS was merged into JAXA in 2000, putting an end to the historically divided structure of the Japanese government’s space programs, mission studies in each of these areas had advanced significantly, leaving ISAS with a number of fleshed-out mission concepts that just needed the go-ahead to proceed to development. The new management quickly discarded the lunar mission concepts as redundant to JAXA’s cooperation with NASA, ESA, and Roscosmos on the Artemis program. Several months of further deliberation followed, with managers, scientists, and engineers debating the various merits of Venus, Mars, and asteroid missions, before a final decision was made to go with the Planet-V concept, a Venus orbiter targeted at atmospheric studies. For Japan’s first mission to another planet, Venus offered the advantages of a less demanding operational environment and shorter flight time than Mars, the destination of the Planet-M Mars orbiter which had been Planet-V’s main competition, much as it had for the United States or the Soviet Union back in the 1960s. And, given that no atmosphere-focused orbiters had visited Venus in nearly twenty years, any mission to the planet, Japanese or not, would clearly have high scientific productivity.

    Following their usual practice, the Japanese announced Planet-V to the world as Planet-“C”, the third in their series of planetary spacecraft (one of the first, part of the Halley armada, had technically been classified as an engineering spacecraft). After the announcement, the Japanese buckled down to begin working on the probe, aiming for a launch date at the next practical window in 2004. As an entirely new design rather than a development of an older vehicle, Planet-C naturally faced more severe design challenges than other missions, and soon enough the program’s managers were having to push hard to have any chance of finishing the mission by the planned launch date. Nevertheless, they pushed, and the spacecraft was completed and launched by the Mu-5 on its third flight in March of 2004 before being successfully injected onto its interplanetary trajectory.

    After launch, JAXA announced the spacecraft’s name: Akatsuki, or “Dawn,” an appropriate name for a spacecraft venturing inwards towards the Sun. Although Akatsuki remained largely dormant during the months-long cruise towards Venus, periodic status checks showed a plague of minor electrical faults, not severe enough to be a serious threat but a concerning sign so early in the mission. Unable to repair the spacecraft, however, mission controllers were forced to watch and wait, hoping that the faults would not worsen or proliferate enough to prevent Akatsuki from fulfilling its mission. As Akatsuki passed behind Venus before Venus orbit injection, anxiety rose to a fever pitch at mission control in Tokyo, only intensified by Fukuro’s ongoing difficulties. When the probe missed its first scheduled communications session after the orbital insertion burn, the control center was very nearly in total despair, fearing a repeat of Fukuro’s propulsion system failure. This time, however, such a failure would most likely mean a total mission loss, as Akatsuki’s attitude control thrusters were simply not powerful or efficient enough to brake it into Venus orbit.

    The day after the first scheduled communications session, however, the Japanese deep-space antenna at Usuda picked up a faint signal from the direction of Venus. Further observations revealed the signature of the spacecraft’s carrier wave, and established that Akatsuki had fallen into safe mode. As reconstructed later by JAXA, it appears that Akatsuki successfully completed its orbital injection burn. Due to a design fault in the main transmission system, however, it had then suffered an electrical fault similar to those it had suffered during cruise as it powered up its transmission system and attempted to realign for its first Earth communication session. As a result, the probe suffered a severe systems failure, though fortunately not serious enough to completely knock out the spacecraft. Instead, it reverted to safe mode, stabilized its orbit, and began screaming for help over its low-gain antenna. Over the next several weeks controllers gradually returned the spacecraft to full functionality, though further electrical faults plagued efforts and attempts to reactivate the high-gain antenna had to be abandoned after the spacecraft repeatedly fell into safe mode while energizing the transmission systems. Nevertheless, they were able to restore the spacecraft to a semblance of full functionality, with all of the instruments functioning normally, and with the spacecraft in its planned elliptical science orbit.

    Unfortunately, the loss of the main high-gain antenna severely curtailed the usefulness of those instruments by drastically limiting their ability to return data, as the low-gain antenna that mission controllers were now being forced to use had only been intended as an emergency engineering backup, and had a data rate of only a few dozen bits per second. It took little work to figure out that at that rate data from the spacecraft’s daily low-altitude passes would each take several days to transmit completely, drastically cutting into its scientific return. The only bright spot was that mission designers had already planned for the spacecraft to buffer and transmit its observational data, so that only relatively minor changes were needed to the spacecraft’s control software to accommodate its new state. Nevertheless, even the limited amount of data that could be returned by Akatsuki was more than enough to make a major scientific impact, with the probe providing significant information about the planet’s upper atmosphere and clouds. Only planned radio science experiments, which relied on the probe’s main antenna, had to be abandoned entirely; attempts to observe and characterize the planet’s powerful and frequent lightning strikes, to track upper-atmosphere circulation, and to study changes and structure in the atmosphere’s fine composition were carried out, largely successfully if on less data and less ability to quickly respond to observations than had been envisioned in the mission’s early design. By the time Akatsuki finally failed in early 2010, after nearly six years of operation, it had firmly established itself as an important milestone in the study of Venus.

    While mission controllers were attempting to solve Akatsuki’s problems, their counterparts elsewhere in the agency were already beginning to work on its successor. While the Planet-V concept had appealed several years earlier due to simpler technical requirements, its Planet-M counterpart had not been abandoned, merely returned to the engineers for technical maturation and concept development. By 2004, it had been refined into a mission laser-focused on imaging the surface of the planet at higher resolution than any previous spacecraft, building on work done by Japan for Earth observation and spy satellites over the past decade. Despite limitations imposed by the lift capacity of the M-V rocket, modest technological development was projected to allow sub-meter optical resolutions, a factor of two improvement over the best available images of Mars from the American Mars Reconnaissance Pioneer satellite. Late in the year, several months after North Korea’s first nuclear test, JAXA received the go-ahead to begin work on what was now Planet-D.

    Despite plans to reuse as much of Akatsuki’s basic design as possible to reduce costs, Japanese engineers were soon finding that they were having to completely redevelop many portions of the probe in order to adapt it to the cooler, dimmer environment of Mars and to avoid the electrical problems that had so heavily impacted the earlier mission, increasing expected mission cost. In an effort to minimize budgetary impacts, Planet-D’s launch was delayed from the originally expected 2005 to 2007, increasing overall costs but reducing yearly expenditures and providing more time to develop and assess the design changes that had accumulated from the older design. Fortunately, this extra time proved well spent after Planet-D’s launch in August and its cruise to Mars, which unlike Akatsuki’s to Venus was uneventful. Shortly after interplanetary injection, Planet-D also received a proper name: Hayabusa, or “peregrine falcon” in Japanese, for that bird’s penetrating eyesight.

    After braking into Mars orbit in late May 2008, nearly a year after launch, Hayabusa began a program of aerobraking much like that carried out by the Mars Reconnaissance Pioneer some fifteen years earlier, dipping briefly into the planet’s upper atmosphere to slowly lower the high point of its orbit. Almost another year passed before it reached its final mapping orbit, only 300 kilometers above the Martian surface, and it began imaging the surface in earnest, gradually building up a high-resolution patchwork of the entire surface. Much as with Mars Reconnaissance Pioneer, this continual observation of the surface has led to the discovery of a considerable amount of seasonal and ephemeral activity, including the observation of several Martian dust devils and avalanches, and, perhaps more prominently, evidence of liquid seeps and flows on present day Mars. Early on during its mission, Hayabusa also discovered several apparently new impact craters that seemed to have exposed fresh water ice to the atmosphere, an important discovery that confirmed the presence of under-surface ice in at least some locations on Mars.

    While Hayabusa was mapping Mars, Japan’s oldest active planetary spacecraft was finally nearing home. After suffering a severe propulsion failure in May 2003, Fukuro had been redirected, with the aid of trajectory planners at NASA’s Jet Propulsion Laboratory, onto a long, looping track that would send it flying past Mars twice to align it to intercept Earth in 2012, all that could be done without the spacecraft’s main rocket engine. One month before impact, JAXA commanded the spacecraft to perform its final maneuver, nearly depleting its remaining attitude-control propellant to ensure that it would hit the ground at Australia’s Woomera Test Range, the vast expanse of Australian outback that Fukuro had targeted since its launch, before ordering it to eject its sample capsule. A quartet of springs held in readiness since launch released, pushing the capsule bearing the precious grains of cometary dust that Fukuro had captured away from the spacecraft and towards the still-distant Earth.

    Fukuro and its sample capsule reached Earth high above the waters of the Antarctic Ocean, carving a hard, bright path against the sky as they bit deep into their home planet’s atmosphere, slowing thousands of kilometers per hour in mere seconds. Before they could endure very much of this, Fukuro shattered, bursting into fragments that rained thinly on the southern coast of the continent, while the sample capsule, armored against the heat and stress of reentry, flew onwards towards Woomera. As it fell to merely supersonic speeds, it stirred for the first time since it had been dispatched by its carrier, ejecting a drogue parachute, then fired its main sail moments later as it slowed beneath the sound barrier, while also triggering a locator beacon. Australian and Japanese recovery helicopters were already in the air as it descended under its canopy, and within an hour of touching down on Australian soil it had already begun its journey to Tokyo’s Institute for Lunar Studies, where the samples would be carefully extracted and analyzed.

    While Fukuro’s sample collector was being dissected, Japanese engineers were already hard at work guiding JAXA’s next planetary science mission towards its destination. With an active Venus mission and a Mars mission in development, Japanese scientists had quickly turned their attention back towards comets and asteroids. Just as Comet Halley had provided the first modest target for traveling beyond Earth orbit, so would these minor planets serve as relatively easy and straightforward targets for another advance in Japanese technology.

    Since the 1980s and the American Kirchhoff and European Piazzi missions, the use of electric rockets, whether ion thrusters, Hall effect thrusters, or thermoelectric thrusters, had become commonplace for Earth-orbiting satellites, replacing older and less efficient maneuvering thrusters. The limited thrust available from electric rockets was not much of a drawback in this application, while the much higher specific impulse they offered compared to conventional cold-gas or monopropellant thrusters meant that satellites using electric rockets could serve much longer than previous designs. Although the Japanese aerospace industry was dominated by domestic concerns, these advantages still applied, and during the 1990s they had begun to develop electric rockets for their satellites as well. This development quickly caught the eye of ISAS, then the Japanese agency responsible for planetary exploration; the high specific impulse of electric rockets would permit their relatively small and payload-limited Mu rockets to lift more capable spacecraft, able to travel to more far-flung destinations or carry more instruments than would otherwise be possible.

    First, however, the technology actually needed to be demonstrated on an operational spacecraft, rather than in a laboratory, or even on a satellite where they would be used intermittently rather than having to constantly fire for years to build up the necessary velocity changes. Comets and asteroids, many of which have elliptical or inclined orbits that are difficult to reach with conventional propulsion, and many of which are located in the inner solar system where temperatures are relatively mild and solar power abundant, were an obvious testbed for this work, and interest had been building in a technology-focused mission aimed at them even before Hayabusa formally started work, though it took several years of study and the freeing up of Hayabusa-related funding before work could start on this experimental spacecraft, named MUSES-B for “Mu Space Engineering Spacecraft” B at the time.

    Such a mission would also offer the opportunity to test other new technologies that could be applied to future missions, such as more advanced computers, new data-transmission equipment, or more autonomous spacecraft control software. The most ambitious of the experiments that gradually accreted onto MUSES-B, however, was also its scientific centerpiece, a small penetrator intended to be fired from the spacecraft as it orbited an asteroid and, as the term “penetrator” implies, penetrate into its outer crust. Penetrators had been proposed for use exploring the Moon, Mars, and minor planets since the 1970s, and in theory had many potential advantages compared to conventional landers for exploring the upper subsurface of those bodies. However, for various reasons none had ever been launched, so that these advantages remained unproven. While small, the Japanese penetrator would at least begin to show whether or not penetrators were actually practical tools of inquiry. Even better, the penetrator could be used to demonstrate one of the newest and least-developed forms of asteroid deflection, kinetic bombardment, where a stream of projectiles would be launched to gradually change the orbit of a threatening body. By actually launching a small projectile into an asteroid, MUSES-B could show the effects such a projectile would have on the target body and experimentally demonstrate the velocity change that could be expected from such an object if it were used to deflect a threatening asteroid or comet. The role of the main spacecraft would be to transport the penetrator to the asteroid and serve as a communications relay between the penetrator and Earth, although it would also carry spectrometers to help extend the penetrator’s precise but localized compositional data to the rest of the body, and a camera for navigation and public relations purposes.

    After more than five years of research and development, MUSES-B was launched aboard an M-V rocket in late 2012, bound for the asteroid Itokawa, which had been discovered only a few years earlier by one of the automated asteroid searches that had been created since the 1990s and renamed after the “father of Japanese rocketry,” Hideo Itokawa, after its selection as the target of MUSES-B. The spacecraft itself was renamed Yumi, or “bow,” while its accompanying penetrator was named Ya, or “arrow,” after the launch, as with usual Japanese practice. Shortly after injection into interplanetary space, Yumi began firing its ion engines, gradually building up speed as it flew towards Itokawa. It took more than two years for it to rendezvous with the asteroid, but earlier this year it finally reached Itokawa, and is currently settling into its final science orbit. Mission controllers say that they are preparing to fire Ya later this year, and are currently debating site selection using Yumi’s images of Itokawa’s surface.

    For the future, JAXA has turned its attention back towards Mars, where it is considering another orbiter mission, or perhaps a small lander or penetrator network. There is even the possibility of cooperation on NASA’s planned Mars Sample Return mission next decade, although that is still some time away and may never come to pass. Nevertheless, the people of Japan still look skywards.
     
    Part IV, Post 24: New space stations for Russia, China and the US
  • Good afternoon, everyone! Workable Goblin's at a conference this week, so I'll be taking back over posting duties. Last week, we took a look at what some of the rising powers of the world were doing with probes. This week, we're looking back at the manned side of things, both for those same nations and around the world. Hope you all enjoy!

    Eyes Turned Skyward, Part IV: Post #24

    The end of the first decade of the twenty-first century found NASA at the center of a frenzy of activity. In space, Freedom and Orion continued their operations, while back on the ground NASA and contractor engineers and technicians pushed forward on the key elements of the Banks Plan. Lockheed-McDonnell was working on the development of the new modules for Freedom’s replacement, building on McDonnell’s legacy with every past American space station, though Boeing’s experience with inflatable modules for the lunar program was called upon for the development of some of the expandable laboratory and habitat modules (which would use a rigid structural core with an annular inflated section [1]). Boeing, for its part, was pushing forward with the detailed design of Saturn II, finalizing details of the vehicles design ahead of a 2011 planned design freeze, and calling in turn on StarLaunch, who served as a subcontractor in the design and development of Saturn II’s landing systems and software. While Boeing had their own ex-Grumman engineer’s X-40 Starcat experience, many of these individuals had spent the last decade working on the Artemis lander program, while StarLaunch’s ex-Grummies were building and operating Thunderbolt. Since Saturn II wouldn’t directly compete with Thunderbolt, StarLaunch was happy to assist--for a price.

    The cash infusement from the Saturn II contract was particularly welcome as it enabled the acceleration of internal efforts aimed at a reusable second stage, a key part of StarLaunch’s original business plan which had been under development for several years, but whose development had been limited by internal cashflow. However, as TransOrbital came online and Thunderbolt’s flight rate rose, the business case for the Thunderbolt L2 increased, and paired with Boeing’s development money, StarLaunch was able to finally unveil the design in 2010. Unlike its potential fully-reusable competition from Lockheed and Europe, the Thunderbolt L2 orbital stage wouldn’t have wings. Instead, it would be a scaled-down version of the Thunderbolt L1--a vertical-landing stage, powered by an RL-10-derived radial aerospike system, which would be used during entry as part of the vehicle’s thermal protection system, actively cooled by residual cryogenic hydrogen. This plug design, SLS’ first internal engine project, was key to achieving effective engine performance from separation and ignition in the upper atmosphere all the way to orbit, and then for the much-lower-thrust final touchdown at sea level.

    While launch vehicle and development on the next-generation space station was underway, Northrop was pushing ahead on scaling TransOrbital’s depot and tech technologies from their currently-operational Centaur form, which saw three payloads placed on the way to GTO in 2010, to the Centaur’s “big brother” Pegasus. Fortunately, the process was relatively smooth, given the technical and construction similarities between the stages. Instead, the main concern between the development efforts and NASA’s 2015 introduction-into-service goal was NASA’s insistence on an actively-cooled depot. While TransOrbital could cope with the minor boiloff of a passively-cooled system, given the relatively small gaps between refills and tug top-offs assured by their planned annual payload throughput, with only a few annual lunar missions NASA wanted an active refrigeration system installed on the Earth-orbiting Pegasus and the Centaur EML-2 depots. The resulting increases in radiator and solar capacity were the main complications for Northrop’s engineers in the system, but with four years the project was well-in-hand, as was the Saturn II--good news for NASA given that the two transport elements were key to their plans for achieving a new station and an expanded lunar base without requiring substantial budget increases.

    However, even these “cost-restrained” American plans dramatically exceeded the opportunities available for some of their international partners. For the Russians, the end of the Soviet era had seen a dramatic reduction in their ability to fund such grandiose plans. Though they began their own reusability program, aiming to recover Neva or Vulkan cores downrange on land, the development budget available was limited, and it only proceeded slowly, corruptly, and with a focus on propaganda value over practical introduction. For the moment, their plans had to be focused on the more practical and near-term: the annual Luna-Pe launches of supplies to the Orion outpost, the replacement of several of the satellites in the aging Mesyat network with a new generation of five Mesyat-II satellites, and the training and coordination of the resulting launch availabilities to the lunar surface they received with the Americans. Foremost beyond this was the replacement of Mir, which after 23 years of service was beginning to rapidly transition from obsolete to decrepit, unlike its well-maintained sister Freedom. The station was forced into retirement and deorbited in 2009. To replace it, Russia had partnered with several commercial investors--primarily from the US--to establish a new station to be partially supported by regular commercial tourist flights, and based on the MOK module core supplemented with TKS-derived labs and temporary modules for commercial orbital science.

    Unfortunately, the station’s plans hinged on the availability of the MOK-2 module, which proved to have been less securely stored over the past two decades than had been initially thought, and problems cropped up persistently throughout the manufacture process. The new “Mir-II” sation remained perpetually poised four years from launch, slipping a year every year. In order to salvage the concept, the station was officially re-designed in 2009 to include a DOS module (which could be built from scratch on available toolings) which would possess nodes at both ends. In the new “final” configuration, the MOK module would dock to one end of this DOS module, while TKS lab, supply and crew modules would dock to the available ports. However, in an “initial” configuration, this DOS would serve as the “service module” for the station, providing basic power, communications, propulsion, and control “until MOK-2 was ready”--a date that remained unspecified even as the launch date of the “temporary” station modules began to move forward, aiming for a launch in 2013. With the downscoping, much of the station’s originally planned scientific value for the Russian program evaporated, but the reduced “initial” station could be developed on the current budget, and would be easier to support on the funds available from tourism flights--and just having the station was enough to satisfy the “soft-power” requirements for the Russian government. Russia and its commercial partners in the West also reached out to other commercial operators, such as capsules proposed for the variety of reusable vehicles emerging around the globe, offering the newly downsized station as a “getaway destination” for tourism or a “commercial lab” for firms interested in space research.

    While the Americans and Russians pushed forward on their own new stations, a third nation had been slowly but surely rising to a level with these two titans of spaceflight: China, Russia’s former partner on Mir. Their close partnership had characterized the nineties and noughties for both nations, with Russian advice and assistance being key to the rapid development of China’s Long March rockets and the Longxing capsule, while Chinese funding had been critical to keeping Mir functioning during the dark days of the early 1990s. Similarly, China’s purchase and conversion of DOS-11 into the Tiangong module for Mir had been an important learning experience for the Chinese, with over a dozen Longxing crew rotation missions ensuring continual habitation by Chinese cosmonauts of the semi-autonomous module, along with ongoing research into a range of topics. However, with the impending retirement of Mir and Tiangong, the Chinese frustrated Russian hopes of continued Chinese funding for Mir-2 in favor of resuming the station plans they had been tentatively drawing up before the collapse of the Soviet Union, this time with real experience behind them.

    This renewed program bore fruit well before the Russian’s own long-delayed Mir-2: the all-Chinese Tianjia-1 module was launched in 2007. Named a hortened form of “Sky Home” which also meant “assembly,” the first station launched since Mir and Freedom was described by Chinese press as an “experimental station” and indeed it bore more resemblance to the Salyut and Skylab programs than its contemporaries. Indeed, limited by the capacity of the Long March 2F rocket to a mass of just 13 tons, Tianjia-1 was smaller than the Tiangong module on Mir. However, it was large enough to verify critical avionics, propulsion, and life support systems, and it was entirely Chinese-built. Over the next several years, a number of Chinese crews visited the station, though it was not continuously occupied. In 2009, after the de-orbit of Mir, a second Tianjia module was docked to a port on the aft end of Tianjia-1, doubling its volume and testing modular assembly, just as Spacelab or Salyut-7 had. Additionally, several Tianjia-derived logistics spacecraft were used to top off propellant and consumables--including one which docked to the aft end of the expansion module, testing transfer of propellant and other consumables across multiple modules.Though such accomplishments for the Americans and Russians lay almost 30 years in the past, the advancements that had taken them almost a decade’s work were replicated by the Chinese only two years apart. Before its retirement and deorbit in 2012, the station served as a proving ground for Chinese engineers to develop the technologies to build and supply their own larger, multi-module station, and they were ready to take the next steps. While their plans were not as grand or advanced as the Americans, the Chinese certainly had more success in meeting their goals than the Russians.

    While the Chinese charted an independent course for developing stations, to the east another nation was seeing their long-term plans finally pay off at Tanegashima Space Center. The pairing of the H-I and the HOPE spaceplane had been the official goal of Japanese spaceflight for almost twenty years, coupling a new reusable spacecraft with an all-Japanese launch vehicle using a high-efficiency hydrogen/oxygen core stage and solid rocket boosters. At this time of the concept’s creation in the nineties, this would have allowed the H-II to boost a higher fraction of its launch mass to orbit than any competing launch vehicle, while HOPE had promised to put them in the lead in the development of reusable orbital vehicles. Unfortunately, the delays in the H-I and particularly the HOPE-C had left Japan’s ambitions standing at the starting line while others raced ahead. However, by the end of the 2000s, Japan had finally begun to make progress on achieving its deferred goals and make revised plans for the future.

    The most immediately visible step forward for the Japanese program came with the maiden launch of the HOPE-C logistics spaceplane to Freedom in 2009. Launching from Tanegashima, the HOPE orbiter made its way to Freedom’s orbit with more than two tons of external cargo nestled in its cargo bay. Upon reaching the station, HOPE employed a radical new method for attaching to the station. A standard Freedom CADS docking ring would have consumed much of the volume available in the relatively small vehicle’s payload bay, while other locations were occupied by the vehicle’s thermal protection systems or engines. Thus, the vehicle was designed to use both of Freedom’s Canadian-built robot arms to carry out a new, alternative attachment maneuver, referred to as “mooring”. In this technique, HOPE would approach the station, then go into free drift. One of the station’s twin arms would then attach to a grapple fixture in the orbiter’s payload bay, holding it fixed relative to the station while the other arm attached to and removed a cargo pallet with the supplies, then replaced it with another pre-loaded for return to Earth (unpressurized downmass being a unique capacity for HOPE). Once the payloads were swapped, both arms could be released, and the spacecraft could return to Japan within mere days.

    Unlike a proper berthing maneuvering, mooring did not create a semi-permanent rigid attachment to the station. While the one-armed mooring was sufficient to hold the relatively lightweight HOPE in place relative to the station during logistics handling, it was impossible to conduct an orbit adjustment burn during moored operations, and the station’s attitude control capacity was necessarily limited. However, in addition to minimizing HOPE’s required orbital life, the “quick-change” pallets for cargo also reduced the mooring time required for HOPE to far less than the docked periods for Aardvark or Minotaur. Additionally, the mooring operation was more flexible, able to be conducted anywhere on the nadir surface where a clear incoming path could be achieved and both arms could reach for the dual-robotic operation station crew nicknamed “arm wrestling”. In practice, the technique worked well, and HOPE completed its mission and returned to Earth carrying a mixed payload of external experiments and failed components for inspection and research without excessive trouble. HOPE’s subsequent flights, and those of its sister orbiter, were major boosts to Japan’s reputation as a spacefaring nation.

    With HOPE in regular operation, however, Japan was faced with the same question facing other nations: what to follow it with. The emerging era of reusable vehicles cast doubt on the long-term viability of the H-II, and Japan risked falling behind the curve as the US, Europe, and even commercial firms invested in partially or even fully-reusable rockets. However, while JAXA’s funding was enough to enable some work, the state of the Japanese economy meant that it lacked either the ability to freely spend on research and development of which Europe was a beneficiary, or the developing commercial market which bolstered NASA’s development capacity. Without the funds to start from a clean sheet, Japan made its goal to develop a reusability solution with the maximum employment of existing development. The result were plans for a new “H-III” program. Based heavily on the H-II, the new system would nevertheless incorporate some major modifications for reusability and cost savings. First, the solid rocket boosters were to be recovered via parachutes and boats for recasting at Tanegashima--something Japan hoped would further reduce the capital cost of the already-cheap solid boosters [2]. Second, they would aim to recover a portion of their first stage. However, unlike the flyback solutions the Americans or Europeans were pursuing, which required extensive weight additions in the form of additional hardware or reserved propellant, Japan would focus on much more limited changes to recover only the primary cost center in the stage: the engines, pumps, and avionics. These would be redesigned into a separate “pod” which could be recovered with wings or a parachute, and which would divide itself after first stage burnout from the expendable (and cheap) tanks. After recovery and return to the launch site, this pod would be refurbished and reused--potentially realizing much of the savings of a full flyback system with a smaller hit to potential payload. In the future, a new upper stage with HOPE-derived thermal protection and wings could be added to create a fully reusable system.

    While their partners and other agencies around the world were developing their own homegrown plans, the Americans were distracted from the preparations for Oasis by a legacy of the past. In the late summer of 2012, public attention was suddenly focused again on NASA, but not for the conversion and transition planning at KSC for briefly supporting both Saturn Multibody and Saturn II, the first suborbital test flights of the vehicle to both land and seaborne-platform recoveries, nor the preparations of the flight articles for the new Pegasus depots and tugs and the new Space Station Discovery. Instead, the agency was rocked on its heels with the sudden end to a part of its history--Neil Armstrong, the first man on the Moon, died in August 2012. It wasn’t the first time the agency and the nation had lost an Apollo astronaut, not even the first since the establishment of Artemis. Alan Shepard had been lost in 1998 to little notice outside of the space community, the first American in space still overshadowed by not being the first American to orbit, and Pete Conrad had perished in a motorcycle accident in 2001 [3]. However, the loss of a name familiar to the majority of the world was different, and his death was mourned widely by the public. The President authorized a state funeral, and an act of Congress in the winter of 2012 formally renamed Shackleton Outpost Orion to “Armstrong Base”.

    Armstrong’s death drew particular comment as a transition point in spaceflight. By this point, even the “token” international astronauts on the Artemis and Orion missions now outnumbered the original Apollo moonwalkers, and the new reusable vehicles were making the Saturns and capsules of Apollo look obsolete. Even as their descendants continued to serve the American program, they were changing with the times, and there were new systems jockeying to potentially replace them. Armstrong's era had passed along with him, and the future was in the hands of active astronauts like Natalie Duncan, former astronauts Don Hunt and Peggy Barnes, companies like Northrop TransOrbital, StarLaunch, and Europaspace, and politicians and ministers around the world.

    [1] Similar to the OTL Transhab/Bigelow designs--the rigid core carries all the structural loads between modules of the station and houses docking apparatuses and the like.

    [2] They’re not going to be rewarded with substantial savings, of course, but ITTL there isn’t quite the lesson of Shuttle regarding reusability and solids. Besides, H-II/III solids are monolithic, and recast at the launch site, which does help operations. At worst, it’s probably about the same price as continuing to expend them.

    [3] He was still riding motorcycles at 69 in OTL, and eventually it was likely to catch up to him.
     
    Part IV, Post 25: Asian space exploration
  • Eyes Turned Skyward, Part IV: Post #25

    By the late 1990s, the caution that had led the People’s Republic of China to reject ambitious plans for expansion into space a decade earlier in favor of a more measured and Earth-centered approach was quickly dissipating. Not only had the Chinese economy continued its rapid growth in the intervening years, giving the Chinese government more resources and a considerably improved manufacturing base to tap for space exploration, but the collapse of the Soviet Union had been a vast windfall, allowing the Chinese to acquire considerable experience in space operations for almost nothing. Between these two factors, the idea of launching Chinese spacecraft to other worlds no longer seemed as impractical or expensive as it had been the last time that the question of greater Chinese involvement in space exploration had arisen.

    Beyond merely financial factors, the irresistible movement of the Artemis program towards the Moon had made starting some type of Chinese program to go beyond low Earth orbit appear to be a national imperative. The immature Chinese human spaceflight program was necessarily still focused on establishing an entirely indigenous space station program like the Americans and Russians had maintained for decades. Chinese manned spaceflight beyond low orbit could not practically be expected for years, but robotic probes offered an immediate response. They were, perhaps, not as exciting as astronauts and cosmonauts walking on the Moon, but nevertheless the Politburo would be able to say that China was out there exploring the universe just like the Americans, and even robotic successes would showcase Chinese developments in science and technology.

    As with Russia and America before, China opted to begin its exploration of space by demonstrating the technology that would be needed for later, more scientifically sophisticated missions. These first missions, although they would carry a few instruments, would mostly be focused on showing that Chinese deep-space navigation techniques, communications stations, injection stages, and other equipment would function properly over the gulfs of time and space that separate the Earth from the other planets. Also as with Russia and America, there were bitter internal debates about where Chinese spacecraft should travel, at least for these early missions. The Moon was nearby and easy to reach, but any Chinese probe sent there would be completely overshadowed by Artemis and follow-up programs. Conversely, Mars and Venus were more distant and offered more complications, but also a field free of any overpowering competition, for good or ill. Given the prestige rationale for the program, the Moon was quickly eliminated from the competition, leaving it merely an argument between Mars and Venus. Like the Japanese at the same time, the Chinese realized that Venus was the easier of the two destinations to reach in both time and energy, and could offer more benign operating conditions once reached provided the toxic and crushing atmosphere was avoided. And, of course, Venus was also the less-explored of the two nearest neighbors to Earth, so even engineering spacecraft could be expected to make discoveries there.

    The public result of this deliberation came in early 2000, when the Chinese announced that they had begun a “Chinese Venus Exploration Program.” This long-term effort to uncover the mysteries of Earth’s closest sister saw the first two orbiters scheduled for launch in 2004. Chinese spokespeople suggested that the technological developments needed for exploring Venus could be applied to “other” exploration efforts, vaguely alluding to possible lunar or Mars exploration efforts. Some American sources picked this up and made a minor furor over Chinese plans to “leapfrog” “Moon-obsessed” NASA by heading for Mars, but they made little impact in the face of demonstrated NASA successes and public disinterest in “racing” China, and the storm, such as it was, soon died down.

    Meanwhile, the Chinese were working hard to pave the way for their orbiters to Venus. Besides designing and building the actual spacecraft, a considerable amount of ground infrastructure and equipment would be needed before the missions could launch. Even before they announced their Venus Exploration Program, the Chinese had begun working on deep-space communications complexes in Xinjiang in the west of the country and on the Shandong peninsula in the east. Together, these two complexes would provide a greater field of view than a single site like the Japanese Usuda complex, but were still far short of allowing continuous communication and tracking like NASA’s Deep Space Network. The Chinese began to angle to build a third center in South America or southern Africa to fill the gap. In the interim, mission operations would be timed so that they could be seen from China if at all possible, and contingency arrangements were made to with the European Space Agency to make use of their deep-space communications facilities if necessary in exchange for other, future considerations. Aside from these communications facilities, the Chinese also constructed several tracking and navigation observatories between the main complexes, creating a great belt of facilities along their country’s centerline and augmenting their power to support distant missions.

    At the same time, they were confronting the challenges of building spacecraft designed to operate not just a few hundred or even a few thousand kilometers above Earth, as Chinese satellites had been doing for years, but tens of millions of millions of kilometers away, so far that commands would take entire minutes to travel from transmission stations on Earth to the spacecraft. More than that, spacecraft near Venus would have to contend with a Sun twice as bright as back home, and with solar wind more than twice as intense, thanks to the lack of a Cytherean magnetic field. Ironically, generating power was another challenge; although the Sun was twice as bright, this meant that solar panels would have to operate at a higher temperature, reducing their efficiency and requiring special design and construction to ensure they provided the required amount of power. Careful design and testing would be needed, and entirely new kinds of testing facilities were built to ensure that the Jinxing, or “golden star” probes, named after their destination, would succeed. Even then, though, there was always the possibility of an unforeseen problem, a simple statistical fluctuation, so the Chinese opted to adopt the strategy that had once been used by the Soviets and Americans in their exploration of space, by sending two spacecraft. Even if one failed, the other might succeed, after all.

    Constructing this network of ground facilities and developing a giant leap in space technology did not come easily, or quickly, and the Chinese were forced to slip the planned launch date twice--first from 2004 to 2005, then from 2005 to 2007. It was also expensive, with the cost of the spacecraft and their ground facilities more than trebling from initial projections in 2000 by the time they were actually launched. Nevertheless, they proceeded, and by mid-2007 their Jinxing 1 and 2 spacecraft were being rolled out to the launch pad atop the Long March 3As that China had designed to launch spacecraft to geostationary orbit, and which were now being repurposed to launch others beyond Earth’s influence. As the launch window opened early that May, Jinxing 1 soared into the skies above the launch pad in Sichuan. Only a few minutes into the flight, however, launch controllers began receiving increasingly worrying telemetry data from the rocket, showing it drastically underperforming after first-stage separation. Performance only worsened as the second stage continued to burn, and within seconds the spacecraft appeared to be drifting irrevocably towards impact in Taiwan. Range safety immediately intervened and triggered the rocket’s destruct charges, sending a hail of debris down into the Taiwan Strait.

    With Jinxing 2’s launch scheduled for just a few days later, mission designers immediately put it on hold plunged into a frenzy of activity intended to identify and cure the failure’s cause as quickly as possible, hopefully before the launch window closed. Careful inspection of launch telemetry and launch vehicle modeling soon showed that the second stage’s underperformance could be entirely explained if the fairing used to protect the probe from atmospheric loads had failed to eject after first stage separation, as planned, but had instead remained attached throughout the second stage burn. Disassembly of Jinxing 2’s fairing and inspection of its parts showed that several of the explosive bolts intended to force the two halves of the fairing apart as part of the separation sequence were faulty; although they appeared to be ordinary explosive bolts externally and to simple tests, they would not detonate on command, apparently due to a bad explosive filling. Later investigation determined that many explosive bolts from the lot that had been used for the two Jinxing fairings had originated in a batch whose whose quality control inspections had been faked, leading to the arrest of several officials at the plant that had manufactured that lot on corruption charges, and the suicide of the plant’s manager.

    In the short run, however, launch managers scurried to replace the fake bolts with real bolts, carefully testing each component for proper function. Just days before the launch window closed, they were able to roll Jinxing 2 out to the launch pad, and it rocketed into the air only hours before they would have been forced to roll it back after a series of last-minuted launch delays. Controllers breathed a sigh of relief when the spacecraft and its trans-planetary injection stage made it to orbit in good condition, and cheered when that stage successfully placed Jinxing 2 on a trans-Venus trajectory a few hours later. Unlike their Japanese counterparts earlier in the decade, Chinese mission controllers saw an uneventful cruise phase over the next several months, ending with Jinxing 2’s insertion into Venus orbit in early November 2007. Over the next several years, it carried out observations of the upper atmosphere, corroborating several results from Japan’s Akatsuki probe in the process. After just over three years of operation in Venus orbit, Jinxing 2 failed in late November of 2010, mysteriously shutting down between its daily communications sessions.

    While Jinxing 2 was quietly studying Venus, work was already well underway on its successors. It and its sister probe had always been intended as pathfinders, more technological experiments than full-fledged spacecraft, and with confirmation that they worked the next phase of the Chinese Venus program began. The next two spacecraft, Jinxing 3 and 4, would take the basic spacecraft design developed for Jinxing and scale it up, creating a platform more than four times as heavy and much more capable. They would also be taking another major technological leap by carrying a synthetic aperture radar, or SAR, to Venus, instead of another crop of atmospheric instruments, allowing them to peer through the all-covering clouds of the planet for a glimpse of her surface. The last spacecraft to carry a radar to Venus had been NASA’s VOIR in the late 1980s, and since then many technological advances had allowed for smaller, light, yet higher-resolution radar instruments that would enable even better maps of the planet’s surface to be made.

    Besides taking advantage of technological advances, the Chinese had another trick up their sleeves to reduce their spacecraft’s weight. Unlike VOIR, which had propulsively put itself into a low, circular orbit, their Jinxing spacecraft would use aerobraking, as had several of NASA’s Mars orbiters since the Mars Pioneer. The technique had been considered but rejected for VOIR due to uncertainties in the Cytherean atmosphere, but improved data from Jinxing 2 and Akatsuki, and the fact that they had two probes, convinced the Chinese that they could afford to risk a problem. Even so, their aerobraking maneuver would be careful and conservative, favoring “doing it right” over “doing it fast”. Without this gamble, Jinxing 3 and 4 would have been much more massive, and impossible to launch on Chinese launch vehicles.

    Initially targeted for launch in 2010, delays in testing their complex radar systems pushed them to the next launch window in 2012. After careful testing of the fairing separation mechanisms, both were lofted towards Venus successfully by their Long March 3B boosters in early March, to reach the planet in August of that year. A long, careful program of aerobraking followed for both spacecraft, taking nearly a year and a half to put them into their operational orbits. Data collection is ongoing, but the Chinese have recently released preliminary maps of parts of the Cytherean surface, on which Doumu, the Queen of Heaven, has joined her sister Hera of the Greeks in naming the planet’s surface features, in accordance with IAU conventions. Many other famous Chinese women and goddesses are also awaiting formal acceptance by the IAU of China’s naming proposals.

    As Jinxing 3 and 4 continue to gather data and build the best map yet of the Cytherean surface, the Chinese are working on Jinxing 5 and 6, which they say are planned to duplicate the balloon mission proposed to NASA as VEIL and implemented by the French and Soviets as Eos, but on a much larger scale. Between them, the two orbiters will deliver a dozen balloons to Venus, then serve as relays while the balloons drift through the atmosphere, tracking wind movement and exploring its deeper reaches. Although ambitious, Chinese mission planners have good reason to hope for success, given their successful history with the planet.

    While the Chinese had been developing Jinxing 3 and 4, their neighbors to the south had begun to contemplate a planetary science program of their own. Although India’s space program is one of the oldest in the world, dating back to the establishment of ISRO in 1969, it has always been focused more on the practical, day-to-day benefits that space technology can provide than the more far-flung and inspirational flights of fancy that most other programs have indulged in, as befits the program of a nation so wracked by poverty as India. Nevertheless, Indian aerospace engineers had always had the same dreams as their counterparts throughout the world, and never stopped thinking about where India could go in space, whether or not it could afford to do so at the time. By the later parts of the 2000s, their efforts began to bear fruit as India developed and became increasingly wealthy and able to look outwards. As with the Chinese a decade earlier, Indian leaders felt that a more vigorous space program would be an effective and relatively inexpensive method of showing off India’s technical prowess, and that a planetary science program would be a particularly visible and yet cheap method of making the space program more vigorous. Thus, even as ISRO began studying an indigenous Indian human space program, they also began traveling along the same well-trod path of analyzing robotic missions beyond Earth orbit.

    Just as with their predecessors, Indian mission planners quickly came to the conclusion that only missions to Venus, Mars, or the near-Earth asteroids and comets could effectively fulfill their geopolitical goals. Mercury, the asteroid belt, and the outer planets would be too difficult, risky, and expensive to reach for a new country just starting to venture beyond Earth orbit, like India, while the Moon was too well-trodden and too heavily occupied by NASA for it to be very attractive a target. Compared to the asteroids, Mars and Venus offered more prominent destinations and a greater link with the cultural zeitgeist, especially in the West. Of the two, Mars offered a somewhat more open playing field: while ongoing JAXA and NASA missions at the Red Planet meant that ISRO’s efforts would inevitably be compared to their more developed capabilities, Venus had shaken off its traditional neglect and its selection would now mean comparisons to Japanese, Chinese, and American missions, with relatively little ability for an Indian mission with a necessarily limited budget to make any significant impact. As the Chinese and Japanese had before them, the Indians chose the more open playing field, electing to begin serious conceptual studies of a Mars orbiter in late 2007. Over the next several years, they continued to refine their mission plans, developed prototype hardware for several instruments, and performed limited tests of instrument and spacecraft hardware while awaiting the government’s decision to proceed with the mission.

    With budgetary estimates, construction timelines, and reliability figures changing from best guesses into firm numbers, that approval came in late 2011, giving ISRO’s engineers two years before the next launch window to build their probe. As with Jinxing 1 and 2, it would primarily be a technology demonstrator, but like those spacecraft it also carried a modest scientific payload, mostly focused on aeronomy, or the study of the upper atmosphere, although a wide-angle camera would also be carried for public relations and to study Martian weather. The last aeronomy-focused spacecraft to visit Mars had been Pioneer Mars thirty years earlier, so there was great opportunity for new discoveries, or at least solid refinements even with the kind of spacecraft India could afford.

    Despite some controversy in the Indian press over the cost of the spacecraft, which although much cheaper than spacecraft from other countries still had a price tag of tens of millions of dollars, work proceeded without interruption from approval until it rode into orbit atop one of ISRO’s new PSLV Mk. III/GSLV launch vehicles in late 2013, with injection into a Mars-crossing orbit following shortly. The success was a much-needed boost for the PSLV Mk. III program, which had seen many problems with the solids that had been added to the PSLV core to boost geosynchronous transfer payload, but it was an even bigger one for the Mangalyaan, or “Mars Vehicle” team, and for ISRO as a whole. Despite competition from larger, more capable, and more expensive spacecraft from Europe, Japan, and the United States, Mangalyaan quickly captured world attention as the “underdog,” the little, cheap craft that was going to explore Mars despite the odds and despite the competition. The long cruise period diminished public interest somewhat, but as Mars Orbit Insertion approached, it began to rise back to what it had been just after launch, and in fact increased even further. The coincidence of a very close cometary flyby occurring soon after orbit insertion doubtlessly accounted for some of this interest, but much of it stemmed from the not altogether deserved reputation of Mars as a “spaceship killer,” a “flying graveyard” that had been eating probes for decades. While it was true that many probes had failed, most of those had been rush jobs from the early days of spaceflight, with little money, time, or experience for quality control. More recent spacecraft had been more successful in proportion to the care and expense lavished on them, and while Mangalyaan was cheap, it was certainly well cared for, and serenely glided into orbit without breaking a sweat. Since then it has been quietly accumulating data on the upper Martian atmosphere, complementing the array of other probes in Martian orbit.

    Even as the Indians were launching their first interplanetary probe, another newcomer was trying their hand at sending spacecraft beyond Earth orbit. Unlike India or China, though, South Korea’s space program could at best be described as “nascent”; Korea had been too small and too poor for a space program to be an affordable proposition for many years. Moreover, unlike the equally poor China and India earlier in their histories, national security tended to tilt them away from a military rocket program which could possibly destabilize relations with China (and North Korea) instead of towards one. Nevertheless, in the last decade of the twentieth century they had started a modest program to develop observation and communications satellites and send a handful of astronauts to Freedom and Mir, a mix of practical applications and public relations-friendly activities, all conducted at low cost. By the late 2000s, their ambitions were growing, fueled by success in their earlier endeavours and reinforced by the boom in partially reusable vehicles overseas, a trend which was making it cheaper and cheaper for them to reach a little farther.

    Of particular note to the Korean Institute for Aerospace Research and the Korean government was the fact that while ambitious probes like the Jinxing series or Japan’s efforts served as proud symbols of technical capability, their launch vehicles--the one element of a space program that Korea lacked--were rarely imbued with the same interest. Few Chinese citizens cared that each and every Jinxing probe had been launched on a Chinese rocket; they only cared that there were Chinese spacecraft out there, exploring Venus. Similarly, while many Koreans were proud of the handful of astronauts that had been launched to Mir and Freedom, few cared that they had hitched rides on American Apollos and Russian TKS spacecraft to get there. The obvious lesson, and one which KIAR assimilated, was that even a modest probe launched by a foreign rocket would be an attractive, and possibly cheap, method of demonstrating Korean technical prowess to the world, just as it had been for many other countries.

    This realization dovetailed with ongoing research at KIAR into so-called “advanced propulsion” methods, particularly their relatively large and active solar sail program. Although solar sails had been relegated to science-fiction and small private efforts in the West, an influential KIAR research, Park Jong-Seok, had developed an interest in them during his education and had pushed for KIAR to develop the technology, arguing that it could allow Korea to “leapfrog” other countries technologically. While the resulting program was small compared to many other KIAR programs, it was still much larger than any effort anywhere else in the world, and by the middle of the decade had developed a variety of prototype sail materials and designs, along with exploring “multi-purpose” solar sails, like sails that would generate electricity in addition to thrust, and other aspects of sail deployment and control. All that was left was to actually deploy a sail in space and show that it would generate the expected thrust, a task admirably suited to a small interplanetary mission.

    Together, these two strands of thought would lead to the beginning of the Korean Interplanetary Satellite program, aimed at sending a modest vehicle similar to the small spritesats proliferating in Earth orbit into interplanetary space, in the process testing solar sail technology. By hitching a ride on one of Northrop’s TransOrbital tug launches and taking advantage of spritesat-standardized hardware, KIAR could launch a spacecraft past the Moon at a cost that made even Mangalyaan look expensive, all while demonstrating a never-before used propulsion technology. iSat, as it became known, went from approval to launch in just under a year, being accelerated to escape velocity by one of Northrop’s tugs in late June of 2011. Solar sail deployment successfully followed, and the spacecraft has been cruising around the inner solar system since, maintaining intermittent contact with its Korean builders whenever it comes close to Earth. KIAR is currently studying follow-up missions that would use such solar sails to travel to Mars, Venus, or a near-Earth object of some kind, though nothing firm has yet been announced. Their success, however, has attracted a wave of study from other programs around the world, who themselves are attracted by the possibility of sending spacecraft beyond Earth orbit. It is possible that, by the end of the decade, every continent except Oceania will have launched at least one interplanetary probe, and it is even possible that the first private interplanetary spacecraft will have been sent beyond Earth orbit. Korea may have been the first to turn skywards because of reusability, but they do not look like they will be the last.
     
    Part IV, Interlude #5: Look Up and See the Future
  • Salutations, everyone! I am the Brainbin, and I am honoured to present to you my final contribution to this wonderful timeline: the last of the pop culture interludes! I must say, it’s been one heck of a ride working on these things for all this time, and I like to think I’ve gained far more than I’ve invested into working on them. With that said, it brings me great pleasure to bring you the following...

    Eyes Turned Skyward, Interlude #5: Look Up and See the Future

    At the 2015 WorldCon being held in Seattle, the gathered assemblage of enthusiasts awaited the beginning of a highly-anticipated roundtable discussion whose panelists included some of the most renowned and acclaimed science-fiction writers, literary critics, and historians of the past several decades. The panel was being moderated by John Scalzi, a novelist and television writer of some renown, on a topic near and dear to his heart, that of how speculative fiction had reflected and interacted with the cultural mosaic of the previous two decades. “Science Fiction and Science Fact” was the name of the panel in question, and the convention’s organizers trumpeted it as an intellectual summit the likes of which were rarely seen. (Other, more cynical observers described it as “so transcendently nerdy, even a convention full of nerds think it’s the nerdiest thing they’ve ever seen”.)

    “Thank you all for coming,” Scalzi said. “It’s really great to be hosting this panel today of all days, what with the latest news from Armstrong Base.” Murmurs of agreement could be heard in the audience, interrupted by one loud, triumphant cry of “ARMSTRONG!” which brought the house down with laughter. Scalzi did his best not to be brought down with it, and continued.

    “A lot of changes have taken place in outer space over these last two decades. Our question is, how much impact have these changes had on the stories we’re writing in the science-fiction genre today, and how much reciprocal impact science-fiction is having on technological developments in the real world. With that said, allow me to introduce the panelists! Would you please give a warm welcome to…”

    He introduced each panelist in turn (with the audience reaction varying wildly, from the polite-bordering-on-tepid applause for obscure academic historian C.A. Baxter to the full-throated cheers for such beloved figures as Ronald D. Moore and Ira Steven Behr). Since Ronald D. Moore - the public figurehead of Star Trek since the 1990s, and undisputed creative authority (apart from the invisible guiding hand at Paramount) since Gregory Garza’s tragically premature demise - was the first to be introduced, he received the first question of the night.

    Star Trek has been with us for a long time now - half a century next year,” Scalzi said. “I know it means a lot to me, as I imagine it means a lot to many of us here tonight.” He paused to allow applause from the audience. “Ever since the 1960s, Star Trek has been credited as having ‘created the future’. Many of us are sitting here right now, playing with our phones, are living proof of that. With that in mind, Ron, how do you and the other Trek writers figure out how to depict future technology, in a world where Treknology is already a reality?”

    “Oooh, great question, John,” Moore replied. “Coming strong out of the gate.”

    “I try,” Scalzi said.

    “Well, it’s a challenge,” Moore said. “I remember reading once, the real challenge isn’t anticipating the need for a car, it’s anticipating the gridlock and urban sprawl that would result from using them. I think the communicorder is the best of example of that. It doesn’t look nearly as advanced to have a single device functioning as a tricorder and a communicator in a world with smartphones.”

    “That goes all the way back to Beyond the Frontier,” Scalzi said.

    “Right, the first Star Trek show I worked on. That really captured the zeitgeist of the 1990s in a lot of ways,” Moore said. “Information technology was revolutionizing every facet of life, even if nobody knew quite what to do with it yet. On TOV, there was an episode where everyone made a huge deal about the computer running the whole ship. Nowadays, computers run everything.”

    “And the setting itself, ‘beyond the frontier’ - leaving the galaxy for the first time.”

    “Right, that was inspired by probes leaving the solar system, and plans for the longer-term lunar landings. There was no easy way to get back to Earth in just a few days. In a way, the Large Magellanic Cloud represented the Moon - close by, but somehow so distant at the same time. All the shots of the Milky Way in the background were inspired by lunar pictures of the Earth. Look how it fills the lunar sky. You’d think you could just reach out and touch it. But you can’t.”

    ---

    One of the emerging themes of Star Trek: Beyond the Frontier was how technological advances could shorten distances between two points. That, coupled with the setting featuring diverse aliens with whom the combined Milky Way crews become acquainted through constant exposure, led many critics to view the series as an allegory for post-Cold War globalization. The Christmas Plot had proven that 1991 was not the “End of History”, but Star Trek’s idea of the future seemed to be a constant repetition of the same tactics which had, in-universe, united the Earth, and then the Federation, and finally the galaxy as a whole. In the long run, surely it could unite the whole universe.

    BTF, as fans inevitably called it, stood in contrast to Exodus, the other major science-fiction series of the 1990s, in maintaining the overall optimism for which Star Trek had always been known. Ronald D. Moore himself was among the few writers who pushed for more interpersonal conflicts within the crew. Showrunner Gregory Garza agreed to this approach, provided that the crew come together to act as a united front against outsiders, rather like a dysfunctional family. Although fans of the show (along with fans of virtually all works of fiction) supported certain characters entering into a romantic coupling, the writers were much more tentative in their overtures in that direction - the franchise had not been terribly successful with past couplings. Captain Kirk and Yeoman Rand in TOV had been a widely-derided joke which ended in scandal; Commander Decker and Lieutenant Ilia were a tiresome and ultimately pointless tease; and Kirk’s romances in the movies had little more substance and longevity than those of James Bond, with him changing love interests in each successive film. Given the enforced long-term proximity of the combined crews, romance was eventually considered a necessity for narrative purposes, not to mention that it allowed the writers to allegorize interracial and same-sex relationships, and directly depict them, in the latter case. (TNV had featured gay characters, but never same-sex relationships.) The most controversial romance on the show was, in a demonstration of societal progress on the issue - neither (meta-fictionally) interracial (both characters were played by Caucasian actors) nor same-sex - it was between Captain Siobhan Ryan and the Klingon Commander Kahv, her First Officer. The relationship was praised for being based on mutual respect and understanding, and of the characters coming to know and like each other over time. Also, the romance remained on the back-burner; each character had far more going on.

    Beyond the Frontier concluded after seven seasons (just like TNV) in 2001, with the crew using a combination of technologies cobbled together from the various alien powers they had encountered and befriended to permanently re-open the wormhole they had inadvertently traversed to the Large Magellanic Cloud. As a result of their tireless efforts, the Federation Council voted to establish a permanent outpost in the region, with which to conduct diplomacy and trade with their new allies - and, in due time, to accept new member worlds. The series ended with Captain Ryan ordering her navigator to set a course through the wormhole.

    The Navigator, a Romulan, had to ask: “Where to, Captain?”

    Ryan merely smiled at this. “Where else? Home.”

    ---

    Beyond the Frontier was a really good experience for me,” Ron Moore concluded. “A great crash course in producing a weekly series. When I started, I barely knew which way was up. And Greg Garza was a great boss - the best I’ve ever had.” The audience dutifully applauded at the heartwarming mention of the late showrunner, still widely mourned by many Trekkies despite his own flaws as a writer.

    “A man I knew personally, and well,” Scalzi said. “And I couldn’t agree more.” He paused, flipping through his note cards. “Of course, Beyond the Frontier wasn’t the only sci-fi show on TV in the ‘90s. Which brings us to the man sitting to your left, Mr. Ira Steven Behr.”

    “Sorry Chris Carter couldn’t make it,” Behr quipped. “Once he heard about the moon base, he went over to NASA to ask if there were any launches to Mars, so he can find the aliens for himself.” The audience laughed, as did Scalzi.

    “He was the man behind Exodus, yes, but you were the man up front,” Scalzi said. “And in its own way, Exodus was even more remote than Beyond the Frontier, despite a much closer setting, in terms of both distance and chronology.”

    Exodus was about personal isolation,” Behr replied. “How you can be alone no matter where you are, or who you are. How it can strike anyone, anywhere.”

    “How do you feel it was influenced by what was happening at the time?”

    “Well, the Christmas Plot played a huge part, of course. A lot of doomsday predictions after that, fire and brimstone. But we’ve always been like that as a culture. The End Times have been around the corner for millennia. Somehow we always survive, and recover. That was another theme of the show. Look at how Europe rebuilt itself after the devastation of World War II. That’s one of the reasons we were always surprised at how everyone described the show as so pessimistic. I can’t speak for Chris, but I found it very idealistic.”

    ---

    Exodus continued to stray from the “hard science-fiction” premise it had debuted with, due to the increased focus on Martian Mysticism by the show’s producer, Chris Carter. However, there was considerable resistance from the other writers, including Behr, as regarded Carter’s plans - or rather “plans” - for the aliens. He didn’t have any, and decided to write many alien plots as open-ended mysteries, leaving the door open for him pick and choose which clues he would use at a later date. Behr, in particular, favoured a more direct approach. “We should tell everyone upfront just what the aliens can do, and why. It’s enough that the characters don’t know.” The compromise that emerged was that the aliens were clearly testing the leaders of the human civilization, in particular the three protagonists played by Tim Matheson, Nana Tucker, and Bill Mumy, for some unknown purpose.

    The only thing which could rival Carter’s love of aliens and mystery plots was his fondness for Judeo-Christian symbolism. To this end, the aliens (whom he personally called “angels”, a term which became widespread on the internet through online “fan chats” with the producers) intended to deliver the humans to the Promised Land (continuing the metaphor of the show’s title) - here represented by an entirely new, Earth-like planet in another solar system. However, the angels would only do this once the humans had passed their “tests”: once Matheson’s character had overcome his insecurities and doubts to become the strong, decisive leader the colony needed; once Tucker’s character had overcome her resentment and hostility to differing opinions and accepted the need for compromise; and once Mumy’s character had overcome his guilt due to his past actions and accepted his ability to redeem himself through good works. This theme of redemption was central to the series finale, in which the entire population of the colony was evacuated to their new homeworld, leaving the solar system behind. The blatant Judeo-Christian themes of the series finale were hugely controversial, generating intense internet discussion (and argument) which would endure for many years to come. Carter and Behr, for their part, both moved onto other projects.

    ---

    “Ultimately, Exodus was an ode to human ingenuity,” Behr said in closing. “And we were inspired by what was happening in space. It juxtaposed the visceral and immediate horrors of a Christmas Plot with the triumphs in the space program at the time. New probes, new missions… so a post-apocalyptic story set in space seemed like a perfect melding of those two opposing ideas. And humans are deeply flawed - that’s what makes them human. But that makes our ultimate triumph, maybe even our inevitable triumph, more satisfying.”

    “Do you think the controversial reaction to the religious elements might have encouraged subsequent writers to focus on less fantastic stories and settings in response?” Scalzi asked.

    “It’s possible,” Behr replied. “After all, we hit on basically every theological precept from Moses to Jesus Christ and back again. Except for resurrection - but then Doctor Who had already done that.” That got a huge laugh from the audience, followed by applause. There was obviously no shortage of Doctor Who fans attending the panel.

    “Dr. Baxter,” Scalzi said, after the buzz had died down. “Since Beyond the Frontier and Exodus had ended at about the same time, what did that mean for science-fiction from that point going forward?”

    “The 2000s were a decade of social activism in support of populist causes,” Baxter explained. “The commercialization of the space industry, and the focus on infrastructure-building, was a double-edged sword. It made space travel more accessible, more mundane, and therefore less fantastic. That coloured people’s impressions of outer space. What so many shows, books, and movies had always treated as commonplace finally was. American producers weren’t able to adjust to this right away - but other creative types elsewhere, who were observing the very same shifts in their cultural perceptions - were able to adapt more quickly to the situation.”

    “Where in particular were these creators?” Scalzi asked.

    “A lot of them were in Japan. Japanese animation was becoming very popular in the United States at this time anyway, and the Japanese approach to just about any topic is to exaggerate its importance to almost comedic excess. Counter-intuitively, treating what was increasingly becoming mundane in the real world as epic and grandiose seemed to be a winning formula.”

    ---

    Space travel and exploration had been a theme in Japanese anime for decades prior to the 2000s, exemplified in such disparate genres as the shoujo action-fantasy Sailor Moon and the shounen mecha Gundam series. These were generally in the vein of most fantasy stories written for younger audiences: focusing on a handful of special, “gifted” individuals. Shows with a space-borne setting which focused on average people with no particular gifts were a key innovation of the twenty-first century.

    The “space trucker”, as Americans might call them, became quite common. Junk collectors and mercenaries with little more than their own jalopy ships and skeleton crews took to outer space, often stopping over at fuel depots - invariably depicted as gas stations. In a way, the depiction of space as something akin to the American West perfectly defined how it had lost the “frontier” aspects of its history while still remaining largely untamed, vast, and empty. [1]

    On the other hand, space becoming more attainable for the average person allowed it to become the ambition of many an underachieving shounen (or occasionally shoujo) protagonist, forced to put his nose to the grindstone just in time to finish at the top of his class and be handed a golden ticket to become an astronaut - a job which required great skill and discipline, despite being increasingly routine - not unlike an accountant or engineer. That said, being an astronaut was still treated as “exciting”, as these series were usually modeled on professional sports animes, where being “the best” at a given sport was the pinnacle achievement.

    Unsurprisingly, given the popularity of anime in the West in the 2000s, these shows would soon reach American audiences. Much like how the jidaigeki films of the mid-20th century had influenced Star Wars, these anime would influence American writers and producers as well, particularly those in the “space trucker” genre, which was already of a piece with the popular cyberpunk genre. It wasn’t surprising that the “space trucker” material was so popular, given how much it owed to the American West, and how trucker culture was so integral to the American concept of “cool”.

    The most ambitious - and ultimately successful - example was Apache, a television show about a small crew of haulers operating out of the titular tramp freighter planet-hopping across the solar system in the (unspecified) near future. [2] This borrowed not only from the convoy aesthetic, but also depicted a romanticized, turn-of-the-(20th)-century “twilight of the old west” setting. The tramp freighter, for example, was outdated, being the proverbial stagecoach amidst a cavalcade of trains and automobiles. The crew of the freighter was, therefore, a crew of romantics, mostly older (though “mature” by Hollywood casting standards - in their 30s and 40s, much like the cast of Alien) and nostalgic for their prime years. The nostalgia was reflected in the eponymous theme song, the original 1960s version of “Apache”, as performed by the Shadows. The production values borrowed from the Used Future aesthetic pioneered by Star Wars, coupled with stories written and set in the American Southwest during the early 20th century, and the visual style of films set in the period (though filmed some decades later) such as Treasure of the Sierra Madre, and the works of John Ford and starring John Wayne. In terms of production values, action-adventure series of the 1960s and 1970s were a clear inspiration, though storytelling techniques and conventions were bleeding-edge modern, with a particular focus on arc-based narrative, taking inspiration from past science-fiction series.

    The owners of the ship were a husband-and-wife team; he had the passion and it was his dream that had brought these well-heeled Terrans to the rugged outer limits of the solar system, whereas she had the brains and the silver tongue, not to mention the bankroll. It had been her money which had bought the ship and the cargo which had allowed them to get the ball rolling on making a living in trading out on the frontier. They were joined by their chief engineer, an old salt (or grizzled prospector, depending on the needs of the plot) who knew and cherished these old ships despite their effective obsolescence. The fourth and final permanent member of the cast was a gifted young woman from Mars who wanted to “see the universe” (despite her great intelligence that could get her a spot at any top Earth university). She joined, as most new characters who upset established ensembles do, in the pilot episode, along with another friend of hers who tragically died to prove that the situation was serious. Thereafter, the cast enjoyed a rotating “fifth chair” position, and the three “core” crewmembers always treated the fourth as if she too might leave at any time, though she never did. Indeed, she proved herself quite valuable as a cook and medic (due to her frontier living experiences) and - since she had studied psychology quite extensively - her role as a de facto morale officer. (Her plans to become a psychologist were the inspiration for her plans to tour the frontier, to see “real people” and not “cast judgements perched atop an ivory tower”.)

    More ambitious was a neo-noir film based on the cyberpunk genre, Spin City. Another retro throwback to films such as Blade Runner, Spin City was set in an orbiting colony in outer space, which was owned by a consortium of corporations (which had paid to build the colony, and whose word was law). It concerned a private detective (naturally) who sought to locate a missing person - only to find him dead, and himself the prime suspect in his murder. On the run from the corporate police, he discovers a vast conspiracy which reaches into the uppermost echelons of the colonial bureaucracy - which, if uncovered, could throw the colony into utter pandemonium. The film became a cult classic, receiving critical praise and several Academy Awards (though only in the technical categories).

    ---

    “Of course, the general trend of Japanese focus on harder science fiction didn’t exclude some very popular exceptions, which had more science-fantasy influences,” Baxter continued. He looked about ready to continue speaking - he had already talked for longer than all of the other panelists combined up to this point - but Scalzi interrupted.

    “Thank you for your interesting findings, Dr. Baxter,” Scalzi said, resisting the urge to hold up a hand in an urge to stop him from speaking. “I think that’s an excellent opportunity to move on to our next panelist, the author of the popular Draconic Instruments Trilogy of urban steampunk fantasy novels-turned-hit HBO series, Judith Rumelt.”

    The precipitous decline of positive responses from the audience while Baxter had been speaking made a dramatic reversal when her name was announced, particularly from a group of cosplayers in top hats and - counterintuitively - tight leather pants. “Judith, your work is set in the 19th century, though it retains many trappings from the present day,” Scalzi said. Why do you feel the ‘present-day past’ setting has become so popular in recent years?”

    “That’s a good question,” Rumelt replied. “People like familiar settings, but they’re aware of how modern technology doesn’t really make for good drama. How many episodes of classic TV shows would be wrapped up in two minutes if the characters had cell phones? There’s a certain appeal to writers of an age where the fastest form of communication was by carrier pigeon.”

    ---

    Perhaps the single defining aspiration of the twentieth century in science-fiction was man’s conquest of the Moon, dating all the way back to A Trip To The Moon, the very first science-fiction film, in 1902. Mid-century futurist exhibits proudly predicted lunar colonies by the year 2000. After the first Artemis mission took place in 1999 and effectively fulfilled this prediction in essence (if not in scope), the sense of anticlimax was palpable. As one of the great science-fiction scribes, Theodore Sturgeon, had so famously written: You may find that having is not so pleasing a thing, after all, as wanting. It is not logical, but it is often true. And so it was in this case. Therein lay the appeal of the steampunk genre, and those which were descended from it: dieselpunk and atompunk were foremost among these; collectively, these genres were known as “retro-futuristic”. This double adjective captured its synergy with nostalgia in general, and with the escapism of more traditional swords-and-sorcerers fantasy.

    In contrast to the epic scope of most traditional fantasy stories (pioneered by Lord of the Rings, which was in turn inspired by epic folktales such as Beowulf), retro-futuristic stories tended to focus on a narrow group of characters, often in an urban setting - which lent itself well to genres such as mystery, suspense, or horror, and a generally dark, gritty, and noir-ish feeling. This was a departure - escapist settings were usually sunnier, more idyllic. Along with a sister genre, the urban fantasy (which featured explicitly supernatural happenings, usually of and pertaining to fantastic creatures), the point of these genres were the juxtaposition of the familiar with the unfamiliar in bizarre, almost surreal ways - and indeed, self-awareness tended to be common. Earnestness didn’t play well in urban fantasy or retro-futurism.

    ---

    “Thank you, Judith,” Scalzi said. “Of course, that’s not the only way to deal with escapist fantasy. For another approach, we have Kathleen Kennedy, Chief Creative Officer of Lucasfilm.”

    Applause rose from the audience, though more tepid than previously, and with a few scattered provocateurs daring to boo and hiss - though never without someone shhh-ing them.

    “Kathleen, how do you feel more traditional science fantasy appeals to today’s audiences in a way that it might not have in the 1970s and 1980s, and vice-versa?” Scalzi asked.

    “I think that nostalgia has had a cumulative effect,” Kennedy replied. “People were nostalgic for the adventure serials that inspired Star Wars, and in addition to that lingering nostalgia, they’re also nostalgic for the Star Wars films themselves. We feel that’s part of the reason why the re-releases and the prequels have been so financially successful.”

    More scattered boos and hisses. Kennedy studiously ignored them - she had a lot of practice. Again, there were no small number of positive responses either. It made for a fine microcosm of fan reactions to Star Wars, ever since the mid-1990s...

    ---

    George Lucas had never been truly satisfied with what would eventually become known as the “original trilogy” of Star Wars films - Star Wars, The Empire Strikes Back, and Return of the Jedi - and as soon as the technology for doing so became available, he vowed to re-edit these films to bring them closer in line with his original vision. He was, at heart, an auteur, who viewed his every concession to the collaborative nature of filmmaking as a personal slight. He took more advantage of the two-decade nostalgia cycle than anyone else (unsurprising given his keen business acumen) when he re-released the trilogy to theatres in 1997, for the 20th anniversary, as part of the Special Edition - with many changes from the original theatrical release, as well as all home video releases up to that point. He would not live to see the fan reaction to these changes, however, given his tragic death in early 1997, when he was fatally wounded as a bystander in a drive-by shooting. [3].

    By this time, pre-production on the first of the prequel trilogy of films had begun. By the term of Lucas’s last will and testament, Steven Spielberg would direct the prequel films. However, he had not completed the script for even the first instalment (Episode I - the original trilogy comprised Episodes IV through VI), and therefore Lawrence Kasdan, who had co-written both Empire and Jedi, was hired on as the principal scribe for the prequel trilogy. Kasdan based his drafts on Lucas’s story ideas, but many of them were found unworkable. Initial plans to make the first movie about Darth Vader’s childhood were scuttled, replaced with plans to focus on an adolescent Vader, about Luke’s age in the first film. Vader, still known by his birth name of Anakin Skywalker, had been orphaned by a traumatic event and left with no alternative but to join the Republican Starforce, where he would eventually encounter a Jedi Knight named Obi-Wan Kenobi, who sensed great power and Force potential in him. The promotional blitz for Star Wars Episode One was unprecedented, with tie-ins covering every conceivable merchandising operation. There was no way that it or any other film could live up to that hype, and indeed it didn’t. Critical reviews were very good, bordering on excellent, but fan reaction was deeply divided. As expected, the film made a mint - nearly half a billion dollars, more than all of the releases of the original Star Wars put together (though still less than Titanic). [4] Though the film won no Academy Awards beyond the obligatory technical nods, it swept that year’s Saturn Awards.

    The two films that followed charted Anakin Skywalker’s seduction by and descent into the Dark Side of the Force, and his romance with Padme Amidala, the mother of his two children (Luke and Leia). In a reversal of fan reaction to the original trilogy, the final film of the prequel trilogy - Rise of the Sith - was the best-received. The actors received consistent plaudits for their performances, particularly the returning Ian McDiarmid as Palpatine. John Williams, naturally, provided the film’s most highly-praised aspect in its original score, with some of his pieces considered on par with his legendary soundtrack for the original trilogy. The visual effects played a large part in bringing CGI out of its awkward, 1990s-era adolescence, though practical effects continued to play a substantial role in the making of the film. The location footage was striking and beautifully shot. The fight sequences were elaborate without feeling unreal or outrageous; Spielberg clearly put his Oscar-winning experience shooting World War II battle sequences to good use. Despite these clear strengths, fans continued to be deeply torn as to how the films should be regarded within the greater mythos - the prequel trilogy had demolished decades of cherished fanon. This, coupled with the poorer box-office receipts of Episode II in particular (the only Star Wars film not to finish at #1 in the year-end box-office tally) would have a chain reaction which affected the franchise’s greatest and longest-lasting rival.

    Star Trek was riding high after three successful “prequel” films of its own by the time of the Star Wars Special Edition. A fourth film was greenlit which would, ultimately, be released in the same year as Star Wars: Episode One. Aware of the need to draw in extra fans, Harve Bennett decided to make the movie about the first mission of the USS Enterprise under Captain James T. Kirk - which would unite most of the characters from TOV and TNV (with the notable exception of Chekov, whose introduction was planned to be delayed until the fifth film). The opening scene featured James T. Kirk being promoted to Captain and assigned command of the USS Enterprise as Captain - now Commodore - Pike handed over command to him. Spock became his First Officer; Dr. McCoy his Chief Medical Officer; Montgomery Scott his Chief Engineer; Lt. Hikaru Sulu his Helmsman, and Lt. Nyota Uhura his Communications Officer. The plot involved the introduction of the nefarious Klingon Commander Kor (as part of the introduction of the Klingons to the prequel films, from which they had been largely absent up to that point). Kor had appeared in one episode of TOV (“Errand of Mercy”) and several episodes of TNV, though never as an adversary of Kirk’s after his initial appearance. This film changed that, borrowing the popular superhero movie approach of taking a character from Kirk’s “rogues gallery”. Kor died at the end of a prolonged space battle sequence, his conspiracy to start a war between the Klingon Empire and the Federation thwarted by the crew of the Enterprise.

    Star Trek: Cold War was well-received by critics, though fans were more lukewarm (as was the case with Star Wars). The film was derided as too derivative; in addition, a vocal minority complained about the continuing absence of Ensign Chekov (who, Bennett explained, was still Cadet Chekov at this point, and would appear in the next film). However, as with all previous Star Trek films, Cold War turned a healthy profit. Thus, a fifth film was ordered, which would premiere in 2001: Star Trek: Eugenics Warriors. As the title implied, the featured villain was Khan Noonien Singh, whose derelict SS Botany Bay was discovered - not by the Enterprise, but by a sister ship, the USS Republic - whose crew was quickly overwhelmed and subjugated by Khan and his men (and women). The task falls to the Enterprise to contain the Republic before it can do any further damage to peace and order within (or beyond) the Federation. Khan was played by an actor also named Khan - Aamir Khan - an actual Indian, as opposed to the Mexican Ricardo Montalban who essayed the role originally (in “Space Seed” on TOV, and then in numerous TNV episodes); this was a deliberate attempt by producers to “rectify” what they described as a “past wrong”. Aamir Khan’s performance as Khan Noonien Singh was intense and brooding, making him a convincing sociopath and supervillain, though he lacked Montalban’s charisma and magnetism, making him less believable as a natural leader.

    Unfortunately, the film opened opposite Star Wars Episode Two, and it lost the ensuing box-office battle. Reviews were mixed-to-negative, and fans - who were very protective of Montalban’s Khan in a way they weren’t even of Kirk, Spock, or Bones - had little nice to say about (Aamir) Khan’s performance, ultimately coming to view the role of Khan as a triumph of talent (Montalban) over authenticity (Aamir Khan). By 2001, the internet was sufficiently large, diverse, and popular that the toxic buzz surrounding the release of Eugenics Warriors would work to help prevent the release of a sixth film in the prequel series. Harve Bennett kicked around a few story ideas, but he and Paramount would officially part ways in 2002 - ostensibly for him to enter retirement, as he was 72 years old - and Paramount itself decided that perhaps the franchise should lay fallow once more (Beyond the Frontier had also ended in 2001), leading Star Trek into its third hiatus in franchise history (after 1969-77, and 1984-91). Bennett had been waiting in the wings from 1984 onward for the resurrection of the franchise, but the Heir Apparent to his informal title as creative head of Star Trek, Gregory Garza, would not live to see the franchise return as a result of his tragic death, leaving it rudderless.

    ---

    “Back around to Ron again,” Scalzi said, flipping through his note cards, “Since that brings us to the present. And the present of Star Trek is, of course, The Enemy Within. Your writing staff had several years to strategize after Beyond the Frontier ended - how did the changes in the space program in the interim affect your plans?”

    “Well, first off, it helps to have a great creative consultant,” Moore replied, and Scalzi (the creative consultant in question) tried his best to look modest. He failed, miserably.

    ---

    “The Enemy Within”, obviously, was the name of a classic TOV episode - the one where Kirk gets split in two (Good!Kirk and Evil!Kirk). The title, while being delightfully evocative, was also an effective allusion: the primary enemy race in this new series, a shape-shifting alien race known amongst themselves as the “Progenitors” but known by the Federation Alliance as “changelings” or “shape-shifters”, obviously due to their ability to take any form, including those of trusted friends and allies. No consistent method of detecting who was an imposter and who was the genuine article was ever discovered - even transporters and communicorders were unable to tell them apart, due to the quantum spectral radiation given off by the Progenitors (who were mechanical in nature). [BC] However, this was only revealed as the seasons wore on. It began on the edge of the Milky Way - the Federation Alliance controlled over 95% of its home galaxy as the pilot movie opened.

    The crew of the lightly-armed “runabout” survey ship, USS Voyager, was exploring a distant spur of a spiral arm on the other side of the galaxy from the Sun when sensors detected a single, very massive vessel from “outside” - beyond the galactic barrier and in the intergalactic medium. The first sign of trouble came when this ship was able to traverse the barrier without any adverse effects - at least, based on the observable evidence. The Captain of the Voyager attempted to hail this unknown vessel, only to receive no immediate response other than being scanned - and then, watching in amazement, for the unknown alien vessel to suddenly (and dramatically) change its shape to match a vessel of the same class as the Voyager. Where no life signs were detected initially, now a few dozen could be found; visual communications were established; a man who vaguely resembled Voyager’s CO (played by the same actor, under heavy prosthetics) was at the helm of this new ship. It was clear that these aliens were mimicking Voyager, but for what purpose? The ship’s communications officer reminded the Captain that many aliens had inscrutable customs; perhaps this alien ship was mimicking Voyager simply to send a greeting. This impression was reinforced by the alien Captain merely parroting everything that Voyager’s captain said to him, and the alien crew gradually looking more and more like the Voyager originals as the conversation between the two Captains continued (achieved through the actor’s prosthetics gradually being removed, with the transitions depicted through CGI “morphing” rather than the more traditional jump cuts). However, the crew of the Voyager soon tired of this and broke off communications. As they discussed their next move, the tactical officer announced that they were under attack - the enemy ship had begun to fire upon them. Fortunately, the enemy’s weapons were pathetic, even in comparison to those of Voyager, who returned fire and was able to defeat the enemy, who self-destructed (in a massive, modulated burst of energy) rather than risk capture. Bemused and totally puzzled, Voyager decided to return to the nearest Federation outpost.

    Where Voyager went, trouble followed - the ship had arrived at the outpost and was under repair when they were hailed by a Starfleet Admiral demanding an explanation: the Voyager had attacked a Klingon freighter in the Garza sector, which was days away even at maximum warp (though still along the outer edge of the galaxy, in the same quadrant). Given the ship’s past itinerary, there was no way it could have been anywhere near the Klingon freighter - which, due to how heavily-armed all Klingon ships were relative to others in their same or equivalent class, had damaged “Voyager” beyond hope for recovery, and it, too, had self-destructed in a similar burst of energy. Since Voyager was still in one piece, it could not have been responsible. Starfleet quickly ascertained that these unusual pulses of modulated energy had to contain encoded signals, though these signals were omnidirectional, making the intended recipient untraceable. All that could be determined was their range; the ship which had attacked the Klingon freighter had to have received the signal from the one which had attacked Voyager. Estimates of this range were confirmed when Starfleet lost contact with Federation research base situated near an interstellar anomaly, about the same distance away from the Klingon freighter that the freighter was from Voyager.

    Voyager had an edge in that they actually had recorded scans of the alien entities as they existed prior to taking on the shape of Federation vessels, and reported this information to Starfleet - in exchange, they would be supporting a task force dispatched to the last known coordinates of the research base in order to re-establish contact. This task force - a full-on Federation Alliance effort - included a Klingon strike force, under the command of Captain Kahv (Jonathan Simmons, co-star of Beyond the Frontier, in the increasingly obligatory “pass-the-baton” appearance). Kahv was depicted as considerably older than he appeared in Beyond the Frontier, far more than only seven years of aging could explain, implying that the show was set some decades in the future from BtF. The task force arrived at the clearly abandoned base - no lifesigns were found, but unusual radiation not unlike the readings taken of the unknown vessel were detected. Captain Kahv suggested beaming a landing party aboard. The Captain of the Voyager demurred, but reluctantly agreed.

    The situation onboard the laboratory was dark and spooky, not unlike a scene out of Alien. All of the known crew members were accounted for - dead, often in particularly gruesome ways. All sensor logs had been wiped; the library computer archives had been accessed and downloaded to some unknown device. Although there was no detectable alien presence, they had nonetheless chosen to send the anti-matter core into an irreversible overload to self-destruct the station - which could not be stopped. The landing party could not remain for long, lest they be blown away as well. The crew were about to return to their ship when a pair of Klingons ominously failed to respond to their team leader - however, when others were sent to look for them, they were quickly found, explaining their lost signal as their communicators having given out, probably as a result of the strange radiation. With nothing more that could be done, they returned to their ship, just in time to see the station explode… and another signal sent out. This was the third explosion, and plotting all three points on a galactic map created a curve, not unlike the circumference of a circle; all three radii neatly converged on the same central point, not too far outside the galactic barrier. The task force decided to head out in search of further answers.

    After traversing the galactic barrier with relative ease (especially compared to the Enterprise back on TOV), the task force almost immediately found what appeared to be a mothership - it was much larger than the ship which Voyager had initially encountered. It was easy to find, given that there was no cover for it beyond the galaxy, an expanse utterly lacking in stars, planets, asteroid fields, nebulae, and gaseous anomalies. However, as the ships approached the mothership, there was suddenly trouble aboard Kahv’s vessel - two of the Klingon bridge officers (the same two who had briefly lost contact with the rest back on the station) suddenly rose up and attacked the others. They were, inexplicably, much stronger than the Klingons, and resistant to their disruptors - but not their physical weapons, which ultimately cut them down, revealing them to be shape-shifting robots - obviously, the very same ones who had encountered Voyager. Kahv, in particular, had a singular moment of triumph when he delivered the killing blow to the second of these changelings, though sadly not before most of his bridge crew had been killed by them. The changelings, though they could not self-destruct as spectacularly as the larger vessels from which they had spawned, still sent signals back to the mothership, which immediately launched “fighters” to attack the task force. The climactic battle of the pilot movie had well and truly begun.

    Although Kahv’s flagship had itself been temporarily disabled (naturally, Kahv and his crew regained control just in time for a dramatic cavalry-style re-entry into the battle) the other Klingon ships were able to fend off the fighters most effectively, and began to advance on the mothership itself. However, it too self-destructed shortly thereafter… and sent out the same radiation as all the other ships had done. Apparently this was not a mothership after all, merely the next step up in a frightening matryoshka pattern. This was confirmed shortly thereafter when, upon returning to Federation space, the Admiral informed them that other such incidents had been reported throughout the galactic periphery, at distances of up to thousands of light-years away. A lingering question remained: what if some of these aliens had been more successful at infiltrating the Federation and its allies?

    Although the pilot movie did not receive a wide theatrical release in the vein of the 1990s Star Trek films, it did receive a limited release in several cities in 2008, such as Los Angeles and New York (making it eligible for the Academy Awards), Washington, London, Paris, Berlin, Tokyo, and Sydney, among others - and, in what was billed as the first extraplanetary film release in human history, on the Orion Lunar Outpost.

    The series proper continued many of the themes and situations established in the pilot movie. The advantage of Voyager’s size was an intimacy the viewing audience had with the small crew - this was a show with no redshirts. As a result, characters did not merely die to prove that the situation was serious - every death had meaning and lasting repercussions. This borrowed more from shows like Apache than from the history of Star Trek, so it was not without some controversy among Trekkies. Furthermore, all of these characters were given story arcs of their own, and the show - even moreso than Beyond the Frontier - was a true ensemble piece. These were also characters lacking in primal, heroic “alpha” traits - strong leadership and decisiveness was less important than cooperation and mutual understanding, which was very much emblematic of the time, and indeed it also reflected how sociologists viewed the “ideal astronaut” - less a John Wayne type, and more an Alan Alda type. (Alda himself was considered to play the Captain before it was decided to cast someone younger.)

    Most episodes were mysteries (often in the TOV mould) or suspense/horror stories. Diplomacy was usually avoided, to make the show stand out from Beyond the Frontier. Intrigue was the show’s primary focus instead - the tone was often somewhere between a noir-ish, hard-boiled detective story and a taut, gritty spy thriller. Voyager was assigned an ongoing mission to seek out and root out possible activity by the alien race which had been called “the Infiltrators”. They could do this because they had first-hand experience with them, and because it was possible for them to operate somewhat clandestinely, lacking the profile (literally and figuratively) of an Enterprise or similar. Details about these aliens were doled out very gradually over the course of the series, but to Ronald D. Moore’s credit, he did sketch out a full backstory for them, and so avoided some of the traps of other, less rigorously-plotted shows (a common complaint regarding Exodus in particular). For one thing, although the Infiltrators continued to attempt to do just what their name implied and attempt to disguise themselves as humans and members of other races for some unknown purpose, enough of them were discovered and interrogated as the seasons wore on that more and more details were revealed about their true nature, including their name: the Progenitors.

    The Progenitors were an example of von Neumann machines, robots capable of self-replication and self-reassembly to suit whatever needs their artificial intelligence deemed worthy. Though popularly referred to as “nanobots” in the jargon of the era, many of the Progenitors (such as their scoutships and motherships).were in fact quite large, consisting of many smaller pieces working together. It was eventually revealed that these alien robots were created millions of years ago, for the explicit purpose of conquering and pacifying alien worlds - by any means necessary. This naturally came to backfire on their creators after their empire became bitterly divided in a civil war - the Progenitor hive-mind could not decide which side to support, and ultimately took a third option - to destroy both sides and choose to govern themselves. However, they did elect to continue their original mission of conquering and pacifying “the enemy”, which led them to occupy their entire home galaxy (Triangulum, as it was consistently known in The Enemy Within) and spreading to other, nearby ones, including Andromeda and (obviously) the Milky Way. It turned out that not only were they facing resistance from the Federation Alliance in the Milky Way, but also from the dominant Kelvan Empire in Andromeda. This news astonished the ship’s First Officer, who was herself a Kelvan from the New Kelva “colony” (now a Federation member world). This would eventually tie into what turned out to be the Federation Alliance’s secret weapon, which comprised the ongoing storyline in the 2014-2015 season. The Progenitors were only capable of traversing the great distance between the galaxies by ship, which took hundreds of years. However, thanks to their experiences in Beyond the Frontier, the Federation Alliance could stabilize wormholes, and eventually Voyager herself found one leading to Andromeda. In the season finale, Voyager travelled through this newly-stabilized wormhole only to find itself in the heart of the Kelvan Empire - and immediately threatened by a massive armada of Kelvan ships, demanding that this strange alien vessel state its business, or face their wrath…

    ---

    The room burst with applause - yet again! - after Moore had dropped his sixth or seventh tantalizing hint regarding the eighth season premiere. Scalzi had been indulgent, but it seemed as though even he was beginning to tire of it. Several members of his panel certainly were; Rumelt was seen rolling her eyes, muttering to Kennedy: “Pfft, Star Trek doesn’t have any shirtless guys in leather pants.” Baxter was openly yawning - ironically, he seemed to find the proceedings even more boring than the audience found him. Behr was pursing his lips so tightly that they had turned white from the pressure. Scalzi noticed this, and decided to put a stop to it. “Thank you, Ron,” he said. “But much as we all love what you’re doing with The Enemy Within, that unfortunately isn’t the subject of this panel.” Moore had the decency to look somewhat sheepish at this. “The question is, what has changed in popular culture with regards to outer space, and why has this been the case?”

    “I think I have an answer for you, John,” Baxter said. Barely-suppressed groans could be heard throughout the audience, but Baxter - an academic, after all - studiously ignored them. “If you’re looking for transitions from space as it was to space as it is today, there’s a singular event to my mind in recent history which symbolizes that change.”

    ---

    And indeed, no single event encapsulated the transition of space in the popular imagination from the “final frontier”, the Wild West writ large, to the “mere” endless expanse of truckers and rangers than the death of perhaps the defining space pioneer: Neil Armstrong, the first man on the Moon. He had lived his life in the decades following his historic moon landing in relative seclusion, eschewing the spotlight as much as possible. Upon his tragic death, there was some debate as to whether he would even receive a state funeral - one had not been bestowed upon anyone who had not been a President of the United States since General Douglas MacArthur in 1964. Armstrong’s would be the first in nearly a half-century. Nevertheless, plans went ahead due in large part to an attempt by the lame-duck President Woods to shore up popular enthusiasm for his administration despite its shaky record on investment in the space program. No doubt Woods was also mindful of his own legacy - many of his critics, who took pains to acknowledge that Armstrong certainly deserved a state funeral, pointed out that he nonetheless would not have wanted one and accused Woods of propagandizing the death of a great man for his own ideological purposes - a chest-thumping exercise of the highest order. [6] Certainly, the touching memorial at the Orion Lunar Base (which was quickly renamed Armstrong Lunar Base in his memory), with the famous photograph of the assembled crew saluting the flag at half-mast during his state funeral, did much to quiet any criticism of the occasion, even among the more cynical of those who had previously done so. Biographies, and indeed the eulogies given at the funeral (which included those of Armstrong’s crewmate Buzz Aldrin, fellow astronaut and former Senator John Glenn, and former President Al Gore) consistently described him as a “reluctant hero” - one who, no doubt, would have welcomed the demystification and democratization of space exploration.

    ---

    Surprisingly, by the time he finished his own eulogy of Armstrong, Baxter had the audience in rapt attention. “It ties into the point Ron made earlier,” Baxter said. “You don’t need to be a Neil Armstrong to go into space anymore. That’s not what’s needed anymore, but it’s because of people like him that it’s possible for people like us to go into space in his stead. In every sense of the word, he was a pioneer… even though he flew aboard Apollo.”

    Baxter did earn some groans at that terrible pun, but also a few chuckles as well, including from Scalzi. “I oughta put that in my next book, Cal - it’ll go along very nicely with the doll pun I’ve already written,” he said. “And a great way to cap some truly great sentiment. Since we’re talking about sentiment, I’d like to close this discussion with some final thoughts from all of you about what you think the current course of space exploration means to people, and what it means to all of you, how your works have responded to it, and will respond, in future.”

    “I think it’s important because it inspires all of us to dream big dreams,” Kennedy replied. “Which is what it has in common with science fiction, I believe. We both want people to aspire to greater than what and where they are now. I’m thrilled with our continuing exploration of the Moon, but I think we can do better than that, and I can’t wait to see it. Onto Mars!”

    Her battle cry was naturally met with cheers from the audience, the first time they had responded to Kennedy in such a fashion all night - clearly taking her aback somewhat. Rumelt, no stranger to controversy, and unafraid of adverse reaction, took the opposite tack.

    “I really don’t think what we’re doing right now is as important as what was done in the past,” Rumelt said. “We’re never going to do anything as big as be the first to land on the Moon. We’re still billions of dollars and decades away from Mars, and all we’ll find there is the same thing we found on the Moon - dirt and dust and rocks. Like I said, I think this is why science fantasy and urban fantasy are so popular. The real future isn’t exciting - it’s boring. A perpetual series of letdowns. The human imagination is so powerful, nothing we accomplish in reality can ever live up to it. When we landed on the Moon, that’s one of the few times we were able to achieve exactly what we dreamed we could. Right now, the questions that science fiction really needs to address are peak oil, overcrowded cities, collapsing biodiversity, crop failure… solving them is a lot harder than just launching a rocket.”

    Rumelt’s analysis met with scattered boos and hisses, but also some applause.

    Behr was next. “I think the situation is a bit more nuanced than that,” he replied. “Technologies derived from space exploration have improved our everyday lives - especially in ways which are relevant to us as media creators. Computing, as an industry, owes most of its early developments to the military in World War II and then NASA in the 1960s. They laid the groundwork for the same technology that allows us to read your books on our e-readers. Solar power technology was perfected to power satellites, space stations, and vehicles - and Earth-based firms were able to use it to power our homes and offices. And how many of us got here with their GPS systems in their cars, or on their satellite phones? None of that would have happened if it wasn’t for the space race in the 1960s. That laid the foundation for our everyday lives, with some gentle guidance from people like Gene Roddenberry, Harve Bennett, Greg Garza, and of course my friend Mr. Ronald D. Moore.”

    Behr’s opinion was met with considerably more enthusiasm than Rumelt’s, and Moore slapped his longtime-rival-turned-friend on the shoulder. “You’re too kind, Mr. Behr,” Moore teased in response. “If only you were this nice to me when Exodus was running.” Behr laughed loudly at this. “To answer your question, John,” Moore continued, “I think we’ve come to realize that the Earth needs to be protected. There’s already been a shift to sustainable development and alternate energy sources - a focus on efficiency and minimalism. We’re also more in touch with ourselves as compassionate, sensitive creatures. We want to change everything for the better - we’re increasingly without taboos, and disdainful of traditions and the way things have always been done. We’ll go to any means to make life better for all of humanity, up to and including changing humanity, if need be. Star Trek has always been hesitant to step into the field of genetic engineering, but I think it’s something we’re going to have a harder time ignoring as we move forward. Everything can be improved, by any means necessary, and anyone who stands in the way of that improvement will suffer the consequences. We’re definitely a more absolutist, less compromise-driven society than we used to be. Ironically because we’re more united than we ever have been in the past.”

    “Well, thanks for that, Ron,” Scalzi said. “A lot to think about - from all of you. I hope you’ll join me in giving a big round of applause to all our panelists.” The audience duly obliged. “They’ll all be signing autographs later this evening, and I’m sure they’d love to see you in the alley. I’ll be there too, and so will I. Thanks so much for coming out, and have a great night!”

    After another round of applause, the crowd gradually dispersed, as did the panel. Baxter muttered his goodbyes to the others before dashing away, leaving the convention hall entirely so he could make the flight back home in time to get a good night’s sleep before his morning lecture. Kennedy also did not linger; she was due to fly back to the Bay Area that evening. Behr and Moore decided to pick up a coffee and chat about old times before heading to their respective booths; only Rumelt chose to embrace her fans at once, as they rushed over to her and gushed about her fiction, which was nothing at all like the fan fiction she had once written. Scalzi observed all of this, silently, thinking that perhaps it would make fine material for his next novel. Writing about science fiction writers may have seemed overly precious and meta for anyone else, but for Scalzi, it was par for the course…

    ===

    [1] Among the OTL anime inspirations for the shows we’re discussing here are Planetes, Space Brothers, Rocket Girls, and Cowboy Bebop.

    [2] No, Apache is not Firefly - Joss Whedon died in the Christmas Plot.

    [3] Because I’ve already got a car full of people due for an accident, and there was no room for George to be a passenger.

    [4] IOTL, The Phantom Menace finished with over $400 million - good for #3 all-time up to that point - though still behind the original Star Wars.

    [5] Essentially, the Borg meets the Changelings - with a dash of the T-1000 from Terminator 2). I’ve also been reliably informed that they borrow from the Reapers in the Mass Effect franchise.

    [6] Neil Armstrong did not have a state funeral IOTL, because his family did not wish for him to have one. I will point out, without further comment, that this family included his widow (and second wife), whom he married in 1994 - a quarter-century after the POD.
     
    Part IV, Post 26: Lunar base planning, commercial moves
  • Eyes Turned Skyward, Part IV: Post #26

    The first half of the 2010s was a busy time for NASA. After all, the last major change in vehicles for the agency had come almost thirty years in the past, nearly beyond the career of any outside of higher management. Indeed, Multibody’s basic design was now older than some of the engineers and technicians working to keep it flying. The last station transition (from Spacelab to Freedom) was not significantly more recent, and even Artemis had been established as a program for nearly twenty years. Even more significantly, NASA was now trying to switch horses in midstream, as it were, substituting Saturn II for Saturn Multibody at the same time it kept up flight rates to support Freedom and Orion and assembled the permanent lunar station at Shackleton. As NASA’s internal team and contractors worked on the projects, the scope of the projects made itself clear in the seemingly endless series of challenges which had to be navigated.

    Even as Saturn II was progressing through its development, the first major payloads it would carry were slowly navigating the path to orbit themselves. The most straightforward of these, ironically, was the foundation of the Oasis-turned-Armstrong architecture: the NASA-contracted, Northrop-built Gateway network, essentially a scaling-up of Northrop’s existing TransOrbital architecture, though with some improvements. The network consisted of three major components. In Earth orbit, there was the Gateway 1 depot, a large depot derived from a Pegasus stage, just as NASA’s cryogenic demonstrator and TransOrbital’s operational depot had been derived from Centaur. As with these depots, the LOX tank of the depot would be made up of a slightly abbreviated Pegasus stage, roughly the size of the LH2 tank in an operational Pegasus, attached to another Pegasus-based tank (this one stretched substantially, to roughly 12 m long) at the front, where it would make up the depot’s LH2 tank. In between the two lay an equipment module, providing insulation between the tanks and housing avionics and docking ports for propellant transfer to and from tugs and tankers. In addition, the Gateway 1 station would also house larger solar panels and radiators than the original TransOrbital depot, which would be used for a new active chilling system designed to effectively eliminate hydrogen boil-off when combined with the existing TransOrbital-heritage passive cooling systems. The Gateway 1 station, when prepared for launch, was just barely short enough to fit within a Saturn-class widebody fairing, and would have the capacity for almost 140 tons of propellant--enough for nearly two complete Pegasus stages. As the Saturn II finished its testing, the final ground-testing of Gateway 1 was completed, and the station was being prepared for launch.

    Even as Gateway 1 was being prepared, other elements of the system were already launching to orbit on the older Saturn Multibody. The first was a pair of Pegasus-T tugs. As with the TransOrbital Centaurs, these tugs were much more limited modifications of the source stage—equipped only with improved autonomous control, multi-layer insulation, and gaseous hydrogen/oxygen maneuvering thrusters, they weren’t intended for long-term propellant storage, only to serve as tankers or to push payloads. Because of this, the first Pegasus-T was able to catch a ride to orbit on a Saturn M02 in 2013, where it began its commissioning ahead of the launch of the Gateway 1 station it would serve. In addition to this tug, another M02 launched the first operational actively-chilled depot, this one a smaller Centaur-based station and a near-copy of the TransOrbital depot, to EML-2 in the same year. There, this Gateway-2 station would serve to “top-off” Pegasus tugs carrying payloads out from LEO to enable the tugs to return back to Earth orbit. Gateway 2’s check-out in EML-2 was sufficient to verify most of the technology changes from the original TransOrbital designs and confirm the readiness of Gateway-1 for launch, which then followed in the summer of 2014. The massive “orbiting gas station” was successfully deployed and tested, and many within NASA breathed a sigh of relief as the critical infrastructure for conveying Armstrong crews to the surface was completed.

    Also entering service were the new generation of reusable vehicles which NASA was relying on to provide the cheap propellant to the Gateway system. Fortunately, here NASA was blessed with a variety of options. In addition to the only partially-reusable Saturn II and the veteran Thunderbolt L1 (both available as a last resort), the agency had contracts placed with both Lockheed-McDonnell and Star Launch Services to supply the network’s thirsty tugs with the hundreds of tons of propellant a year that the system would require using their coming generation of fully-reusable vehicles. Indeed, for both companies, this contract was key to closing the business case for the development of their reusable vehicles--NASA’s requirements provided each a solid base of flights to supply Armstrong’s propellant needs. The result was that the early 2010s saw not just StarLaunch’s development of the L2 orbiter for Thunderbolt but Lockheed’s more exotic two-stage Starclipper spaceplane. It was an all-out race between the two to see which would claim the honor of the first operational fully-reusable launch system.

    Although SLS had been operating the reusable first stage of Thunderbolt for more than five years, Lockheed-McDonnell initially held a decisive lead on Starclipper’s development, benefitting from their myriad of other military and commercial aerospace ventures and resulting deep pockets. While Paul Allen was a billionaire, even his wealth seemed insufficient to self-fund Thunderbolt L2 development, especially given his reluctance to sink his entire fortune into the launch business. However, the approval of Saturn-II and resulting granting of a sub-contract for the landing system from Boeing to SLS was a major boon for the firm—not only were the funds themselves appreciated, but Allen was able to leverage them into a major second round of investment capital and push Thunderbolt’s L2 second stage development, led by Don Hunt, to full throttle. In spite of this, Starclipper’s two years of formal development and years of studies provided quite a lead to make up. In 2012, the new scaled-up X-33 which would serve as Starclipper’s first stage made its first development flights from Edwards Air Force Base, the same place where the original X-33 had been launched years before. Given this past history, there was less to verify with regards to the vehicle’s aerodynamics and overall design, and instead the test program that year focused primarily on the actual functionality of the booster’s engines, tanks, avionics, and thermal protection system. In a near-mirror of the flights made just over a decade earlier by its little brother, the maiden Starclipper booster North Star passed its acceptance tests with flying colors, including a long-range glide test to Malmstrom AFB—not just a verification of the vehicle’s glide-forward distance but a proof-of-concept for national-defense polar orbital launches. Following this flight, the vehicle was transported by aircraft to its new operating base at the ALS Matagorda Bay launch site, where it began a series of additional shakedown flights to the new Clarence “Kelly” Johnson landing site in southwest Florida. With the first booster operational, work proceeded on readying two others—one for national defense launches, the other as a “ready backup”.

    While work proceeded relatively smoothly on Starclipper’s booster, the orbiter portion of the system proved more of a handful to the engineering teams. Faced with the challenges of achieving a lightweight vehicle capable of serving as a second stage, reaching orbit with a meaningful payload, then returning to Earth safely, it was perhaps no surprise that its development was complex and expensive. While the booster was making its first test flights, the testing of structural pathfinders and aerodynamic mockups continued throughout 2012. It wasn’t until the summer of 2013 that the first captive-carry and glide tests of the Starclipper orbiter were conducted from the back of a modified Lockheed Bistar freighter. However, once the flight testing began, the pace was rapid, beginning with a series of captive-carry flights. The test campaign concluded in a dramatic fashion with a pair of “approach and landing” tests, where the orbiter was released in mid-air to glide free of the carrier plane and into a runway landing to prove the orbiter’s approach and landing systems. With the testing concluded on each portion of the system, Starclipper was ready for its maiden flight in April 2014.

    However, while Starclipper’s testing had been underway, StarLaunch had been closing the gap. Given that they had their own first stage already prepared, their development tasks were simpler, and the barrel-shaped Thunderbolt L2 stage was significantly less complex both structurally and aerodynamically than Starclipper’s spaceplane. Thus, around the same time that their Lockheed-McDonnell competition was preparing Leander for her first powered flight, Don Hunt’s SLS team was shipping the first Thunderbolt L2 stage to Wallops for flight testing. With its vertical takeoff, vertical landing design, SLS’ Thunderbolt orbiter benefited from being more easily tested separately from its booster, roughly mirroring the test hops which the Thunderbolt L1 booster had conducted in 2004. In these flights, the vehicle’s systems were shaken down and tested, with a primary focus on its avionics, the RL-10-based aerospike cluster, and the vehicle’s base-first thermal protection system. Thus, though the Starclipper team had almost a three-year head start, by the time of Starclipper’s first operational orbital flight in the summer of 2014, Thunderbolt was trailing by bare months, planning their own orbital demonstration in the late fall.

    Both vehicles were faced with an immediate demand for their services. While they had been completing their tests, the Gateway-1 depot had been launched to orbit in July 2014, with filling immediately beginning using Thunderbolt L1 and Delta 5000 launches while NASA awaited the operational debut of the Commercial Fuel Services (CFS) vehicles. In 2015, the two competitors quickly began to ramp up their flight rates as they settled into routine operations not just supporting Gateway-1 but also sending satellites up to TransOrbital’s second generation Centaur tug and depot based on the Gateway-2 design for transport to GTO. However, the Americans weren’t the only ones seeing their plans for reusability paying off—the Aetos first stage of EuropaSpace’s Aquila system made its first taxi tests in August 2014, on track for ESA’s own system making its debut into service in 2015 or 2016. This was also key to NASA’s plans, given that the modular design of Discovery (whose modules were proceeding from design into assembly and launch preparation) depended in a large part on Aquila’s ability to launch and return large-diameter payloads of up to 30 tons in a single flight—almost triple that of Starclipper and five times that of Thunderbolt L2. In addition, Italian engineers were making progress on the design of ESA’s own depot system and their aerobraking interorbital tug, Prometheus. This coming ability to contribute to the propellant supply for the Gateway network was a major component of Europe’s barter for ongoing flights of its astronauts to Armstrong, Freedom, and (once it was launched) Discovery.

    While the Americans and Europeans were achieving successes in advancing the next generation of space access, their Russian equivalents were more worried with securing their success in the previous generation. With the retirement and deorbit of the decrepit Mir in 2009, the Russians had been left without a space station on orbit, leaving them arguably less capable in spaceflight than even their Chinese former partners, thanks to the latter’s Tianjia program. While their cosmonaut’s flights on American Orion expeditions to the Moon was some comfort, it wasn’t much for a country that had grown used to being one of the two dominant leaders in spaceflight. Much of the weight of restoring the glory of the Russian program (and of funding any advancements to restore parity with even the Chinese or Europeans) depended on the plans for the semi-commercial Mir-II station. Following its “re-scoping” in 2009, in which the MOK-2 module originally built in the 1980s for Mir I was re-designated as a “phase two” addition to the station, work had proceeded with the preparation of the new-build DOS “service module” which had replaced MOK-2 as the central core of the station, and the TKS subsidiary modules had already been largely completed. However, the station wasn’t yet entirely free of the delays which had haunted it for half a decade. While the original plan as re-conceived in 2009 would have seen the station’s modules begin launching in 2013, the DOS-SM was only barely shipped to Baikonur towards the end of the year, and the station’s launch would slip one last time into 2014.

    However, this delay would indeed prove to be the last. In February 2014, as the Americans were readying Armstrong, Gateway, and Discovery, Mir-II finally got off the ground as a Vulkan carried the DOS-SM to orbit. Later in February, the first crew of cosmonauts would take up residence in the module, and oversaw the addition of the first two subsidiary TKS labs necessary to support the “interim” crew capacity of 6. With this complete, the station was visited in October by its first paying visitors when the operations crew of three was joined by another TKS module flown by one cosmonaut and carrying two space tourists. With Mir-II finally in orbit and earning money, Roscosmos was able to finally begin thinking about preparing MOK-2 to join the station—a task made easier by the ability to rely on DOS-SM for much of the command, control, and crew life support functions which MOK-2 would have hosted in the original Mir-II designs. In addition, at long last, the Russians were able to begin the job of conceptualizing their own reusable launch system to replace their Vulkan and Neva launch systems, just as Saturn II, Starclipper, Thunderbolt, and Aquila were replacing other launch systems of the same vintage.

    While the Russians were finally following through on their plans for their new station, the Americans were coming closer to fruition of their own plans for the expansion of the Orion into its new guise as the permanently-staffed Armstrong Base. While each Orion mission was as capable a platform for scientific exploration as the entire Artemis mission series (much as each Artemis flight nearly matched the sum total of the Apollo missions), they were still limited in their ability to expand the base and explore for extended durations beyond a hundred kilometers around the South Pole. Additionally, while the base’s facilities--a three-story expanded Artemis habitat, an Artemis-derived logistics lander, and two pressurized rovers--were more than sufficient for the early month-long missions, they were distinctly cozy for the longer 3 and 4-month flights that had followed in the early 2010s, particularly as the habitat’s volume (and even the volume of the logistics lander freed by consumables usage) was becoming crowded with longer-running experiments and new gear transported up by the annual flights. Finally, the outpost’s two main modules were securely bolted atop their descent stages, making transporting or linking them with any expansion modules impossible.

    Armstrong was designed to improve on its predecessor in several ways. Some of these were in the transportation architecture; in addition to being cheaper, the topping off of the descent stage’s propellant and the assistance of the Pegasus tug would enable landing 20-ton modules on the lunar surface, an increase from the 14.5 tons of Artemis and Orion’s single-H03 cargo missions, and even on the 17-ton payload of the dual-launch crew landings. This use of the Gateway network and reusability was key to the plans for maintaining a permanent staff at Shackleton crater. However, Armstrong also represented a switch to a more permanent, expandable base design as well as a more sustainable access architecture and a slight increase in payload capability. First, the base for the first time would detach landed modules from their habitats and place them on the surface--a change which would serve to dramatically ease operations around the base, preventing a recurrence of the famous “Little fall” by Artemis 7’s commander. However, the change would also make it possible to connect the base’s modules directly, in a surface-bound version of the modular assembly used in every space station since Spacelab in 1978. To accomplish this, the base would use a new rover design, called the All-Terrain Lunar Activity System (ATLAS). This was a two-part rover, each part having a tripod of long wheeled legs. These halves could extend up on opposite sides of a descent stage, grab a payload from the top, and then move it off and lower it to the ground. Once at ground level, ATLAS could then roll the modules around to connect them. Once linked, permanent legs could deploy from the modules to hold their position and level on the ground, and ATLAS could detach to other work.

    This crane/rover capacity figured heavily in the design and assembly sequence of Armstrong. The plan called for three cargo landers to deploy several major modules. The three primary modules were based on the same vertically-oriented 5-meter diameter modules used in the Artemis, assembled in a triangle. Each would have two levels within its rigid portion, plus an inflatable dome--equivalent in size to the Orion habitat module. The modules could be linked to each other as well as other module by new “Surface Attachment System” docking ports. SAS was a port standard similar to CADS or LPAS in concept, but with a taller, more traditionally “door-shaped” opening. The first module to land would be fitted out as the “operations and habitat core,” essentially a lunar-bound equivalent of Freedom’s Habitat and Service Module Challenger. In this role, the operations would house the base computers, main life support systems, and primary crew-support functions including working spaces, the galley, and a hygiene station. In addition, the module would launch to the lunar surface with the base’s main airlock attached at one of its side ports and carrying the ATLAS rover. The second cargo lander would carry the station’s science and lab module. The lower floor would be entirely dedicated to a geology lab and EVA support, including the suitports and an SAS port initially housing a backup airlock, while the upper level would be devoted to biological and physics experiments. The third module would be devoted initially to cargo and logistics, with heavier equipment and spares on the lower level and food and other supplies on the upper story, but it also included reserve suitports, life support, and a second hygiene station, and its internal volume was intended to be repurposed as the supplies were consumed, such as for expanded living quarters or equipment storage.

    In addition to these three main modules, there were several smaller modules planned. The most notable was a new pair of pressurized rovers, much like the design already in service at Orion. This “camper” had seen heavy use during Orion, and thus the decision had been made to provide a new pair of rovers, which would include the new SAS ports among other improvements to enable longer-duration traverses. This would not only allow the rovers to dock directly to the base and avoid the necessity of an EVA to transfer crew and supplies, but would also allow the two rovers to dock to each other in the field during long traverses, providing additional contingency options for trips which might be as much as a week’s drive from Armstrong proper. One of these new rovers would arrive on each of the second and third cargo landings, atop the science and logistics modules. The other secondary module was an experimental “semi-rigid” module. This would consist of a deployable floor frame with an SAS port in a single vertical wall. The sidewall of the rest of the sausage-shaped module would then inflate, like the habitat domes on Artemis and Orion and the new modules under design for Discovery, with the rigid frame serving as the basic structure and an attachment point to the rest of the base. The module’s main purpose would be to test semi-rigid expandable modules for surface bases, but it was intended to be attached to the science module for expanded lab space (potentially including a small greenhouse if fitted with appropriate lights) and could also be buried under regolith to serve as an improved “storm shelter”. Combined, the modules of Armstrong would offer over two thousands square feet, making the base “house-sized,” as the Public Affairs Office insistently noted.

    While Armstrong’s modules were being assembled by Boeing in Bethpage ahead of the planned 2015 first landings, NASA was finishing operations at Orion. This was driven by a combination of factors, ranging from minor contributors like the need for experienced lunar mission planners’ input into the development of Armstrong’s hardware to the more serious, like the minimization of tricore Saturn Multibody H03 launches during the transition to Saturn II. However, the most dominant was the depletion of the outpost’s pre-emplaced supplies. Originally, Orion had landed with roughly 12 crew-months of consumables, and each Orion crew added only another three weeks between their lander and the annual Luna-Pe resupply vehicles, while lasting six weeks to three months. This meant that over time the stocks of consumables at the base were being steadily depleted, though this had been anticipated during initial planning. Since the plans called for Armstrong’s maiden crew to rely on Orion’s habitat as a “construction shack” during initial fitting out of the base, it was necessary to leave a supply reserve, which could then be added to Armstrong’s initial supply cushion if some part were not consumed. Thus, the final Orion mission would be Orion 5 in 2012, lead by Aaron Altman, a veteran of Artemis 8. During their three months at the outpost, Altman’s crew primarily focused on closing out operations at Orion, finishing and collecting data and samples from a variety of long-running experiments and preparing the site and the modules for two years of inactivity and remote operation before Armstrong’s modules would begin their arrivals.

    The final hurdles to be cleared before the beginning of Armstrong and Discovery’s operations was the introduction and testing of the new Saturn II first stage, upon which the cheap launches of the base and station components and crewmembers would depend. Fortunately, Boeing was an experienced astronautics firm, and both they and their subcontractor Starlaunch had experience with the introduction and testing of reusable vehicles--Boeing with the heritage of Grumman’s X-40 Starcat, and Starlaunch with the more recent Thunderbolt L-1 and their ongoing work on the new L-2. Thanks to this, work had been proceeding with remarkable dispatch since the program’s official approval in 2009, and even as the Michoud Assembly Facility continued to roll Saturn Common Cores and Boosters for Freedom and Orion operations, the first structural test and flight Saturn II first stages had taken shape beginning in 2011. In 2012, the first integrated core had been shipped to Stennis to undergo rigorous qualification testing of the new engine cluster and the core’s structural design. Later in the year, while S-2-B001 was still in testing at Stennis, S-2-B002 was shipped by barge to Cape Canaveral to begin facilities checkout.

    With the launch of Orion 5 in July 2012, the flight rate at KSC saw a noticeable drop off, enabling the dedication of a VAB cell and Mobile Launch Platform to preparations for Saturn II. Also undergoing preparations was the new Saturn Landing Facility (SLF) which had been constructed on the grounds of the decommissioned LC-13. Preparations at LC-13 such as the new Booster Preparation Facility (BPF) and the landing pad itself were given priority, and S-2-B001 soon arrived fresh from its testing in Stennis in early 2013 as S-2-B002 was belatedly shipped to Stennis for its own qualification for flight. As with Starcat and Thunderbolt before it, Saturn II made its first flights in short hops from deployed landing gear at the SLF, verifying the ability to conduct a safe landing. After a series of escalating altitude hops concluded in mid-2013, Saturn II made its first flight from KSC in July, launching from LC-39B in a suborbital flight, flying downrange almost 50 kilometers, and then returning to a successful landing at the SLF. A repeat flight was conducted in late August, and the first full Saturn II demonstration flight took place in November, with the S-IV upper stage delivering its mass simulator to orbit at roughly the same time that S-2-B003 touched down at the SLF. In 2014, Saturn II was officially qualified as operational, at least in the “Medium” configuration, and the flight rotation of six boosters took up the slack with hardly a hiccup as Saturn II officially took over the “milk-run” flights to Freedom, even as testing continued on landings on the downrange recovery of cores and of the Saturn-II Heavy configuration, which debuted in November 2014.

    With the testing of the first Saturn II Heavy, the final roadblocks in the preparation of Armstrong were cleared: Gateway was up and running, and Armstrong’s modules were prepared for flight. Over the course of 2015, the first three operational Saturn-II Heavies lofted their payloads to LEO, handing off the landers carrying the Operations, Science, and Logistics to the Pegasus tugs which would convey them to EML-2, from which the landers delivered the three payloads safely to the lunar surface. Today, all that remains is for the arrival of the first Armstrong base crew to follow the base’s hardware to the surface and complete the assembly and commissioning of the first permanent moonbase, a flight scheduled for early spring of 2016. Meanwhile, within NASA, attention is shifting to the plans for the launch of the first modules of Space Station Discovery in 2017, and the subsequent retirement and de-orbit of Space Station Freedom. With the commissioning of a new wave of reusable vehicles, a new generation of space stations, and the first permanent moonbase, a new era is dawning in spaceflight--one that holds the promise of accomplishing long-delayed dreams dating back to the era of Apollo.
     
    Part IV, FINALE: The Future
  • Eyes Turned Skyward, Part IV: Finale

    In the late 60s, at the height of the Apollo program, NASA had conceived of a grandiose vision for the future, the “Integrated Program Plan”. This plan, in several variations, had amounted to a wish-list for the agency’s future, based on the assumption of a budget continuing at or above that of the Apollo-era peak. In these visions, reusable spaceplanes and large Saturn boosters would have been used to access Earth orbit, supporting a fleet of nuclear-powered orbital tugs, space stations, and depots, which would have enabled ongoing exploration of the Moon and even missions to Mars. The drawdown of funding in the wake of Apollo 11 was a hard reality check for these ideas, and under Administrator George M. Low, the agency had instead followed a cheaper program focused on just one aspect of the integrated plan, the orbital space stations--a decision that lead directly to Skylab, Spacelab, and Space Station Freedom. However, less than half a century after that grand vision, Low’s policy of a largely conservative, incremental development of capability combined with the unanticipated explosion of private spaceflight has seen many of the lofty hopes of the Integrated Program Plan become a reality. Through incremental development, ongoing advocacy, and more than a little luck, the agency that started with just one leg of the Integrated Program Plan has achieved almost all of what those planners envisioned and more.

    In the late 60s, perhaps the most anticipated element of the Integrated Program Plan was one that was common in almost every piece of science fiction--the reusable space shuttle. In the Integrated Program Plan, such a vehicle would loft satellites, space tugs, propellant, space station modules, logistics cargo, and even crew to orbit cheaply and regularly before returning to its launch site for another mission. At the time, the potential for such a “space shuttle” had made its continuation, rather than that of the space station program, a strong contender for Low’s recommendation as the primary direction of the agency if only one path was to be followed. However, skepticism about the cost of technology development and the necessary flight rates lead to pressure from the OMB and White House to seek more cost-effective development paths. The dream of a fully reusable, multi-purpose space shuttle, though, is one that never truly died. With the Star Launch Thunderbolt L2 system joining the Lockheed Starclipper and the ongoing testing of the European Aquila, fully reusable launchers are driving the cost of access to space below what would have been wild dreams just a few short decades prior. Supplemented by NASA’s next-generation partially-reusable Saturn II, the ability of these reusable vehicles to launch spacecraft, cargo, logistics, and propellant is key to the current boom in spacecraft launches and the ongoing preparations and support of Armstrong Base at Shackleton Crater and the launch of Space Station Discovery. Given the necessary flight rates from launch pads in Matagorda, Wallops, and Florida, the sight of one of these reusable “shuttles” rising on a tail of fire can be seen at least monthly from nearly anywhere on the East or Gulf coasts.

    In orbit, the Integrated Program Plan envisioned at least one large “Space Base,” augmented as necessary by smaller platforms in other orbits, as well as an assortment of depots and tugs. This was the main element of the Plan’s vision that survived cancellation in the 60s, but even though Space Station Discovery and the Gateway depots at LEO and EML-2 may be seen as just a continuation of of this legacy, there are differences. Unlike its predecessor Freedom, Discovery isn’t designed to ever truly be “complete”. Instead, it consists of a core group of modules for supporting the station’s ten to fifteen-strong crew and a variety of outlying lab modules. With the European Aquila and Lockheed Starclipper, the proposal is that these subsidiary modules can be returned from orbit as necessary for refurbishment or reconfiguration to support new scientific objectives of the station before being relaunched. The creators of the Integrated Plan also likely would not have anticipated the variety of origins of the various other stations circling the planet, from the partially commercially operated Mir-II to the Chinese Tianjia program. The commercial heritage of the Northrop Centaur-derived TransOrbital depot and its cousins, the Pegasus-derived Gateway 1 in LEO and the Centaur-derived Gateway 2 at EML-2 would also be surprising, as might the flights of orbital tourists to Mir-II and (in some proposals) on specially-fitted “L3” Thunderbolts or Starclipper shuttles. In the era of the IPP, that anyone but NASA would lead in space infrastructure development or crewed flights would have been unimaginable, where today it is commonplace for NASA to rely on commercially-developed vehicles and technologies for its space launch needs. While Saturn and Apollo are still core elements of space exploration, they too have changed with the times.

    One major element of the Integrated Program Plan of the 1960s that has not materialized and seems unlikely to do so is the “nuclear shuttle”--a reusable nuclear thermal transfer stage which the IPP envisioned being used to transfer payloads between Earth orbit, geosynchronous orbit, the Moon, and beyond. While nuclear power is slowly recovering from damage done to its reputation in the 70s on the ground, concerns over launching even RTGs to orbit make the prospect of a full-scale orbital reactor dim, much less a nuclear rocket. However, the IPP’s nuclear shuttle does have a direct analogue: its role as an interorbital tug has fallen to the Northrop Centaur, NASA Pegasus, and the planned ESA Prometheus cryogenic tugs. Though less efficient than their nuclear-powered inspiration, once paired with the Gateway and TransOrbital depot network, this fleet of more than half a dozen orbiting tugs is capable of easily moving propellant, cargo, and even crew around LEO and cislunar space, and even to Earth escape.

    Of course, one of the key tasks for this tug network is the support of crew, logistics, and cargo flights to the new Armstrong Base at Shackleton Crater. As the Oasis program sees Orion’s “soonbase” grow into a fully-operational outpost permanently crewed by four astronauts, the depot and tug network has already been hard at work transferring the base modules launched by Saturn II as far as lunar orbit, and crew are soon to follow. The house-sized trio of main modules, as well as the experimental semi-buried inflatable greenhouse and new rovers, represent a true outpost on humanity’s nearest neighbor. From the base at Shackleton, crews are planned to stage traverses of as much as 500 kilometers around the poles, exploring nearby craters and mountains, and conducting a wide variety of scientific investigations including preliminary small-scale experiments with excavating and extracting water ice for the base’s own use and possible electrolysis into propellant. In addition, Shackleton already plays host to the Lunar Infrared Tracking Telescope, the FROST-II radio telescopes, and the growing scattering of omnidirectional surface-mounted antennas that make up the Lunar Low Frequency Observatory. While the four-person outpost may not currently live up in scale to the dreams of some in the 60s, Armstrong certainly lives up to the vision of a lunar base, and seems likely to continue to expand in the coming decades.

    It is certainly in the level of beyond-Earth exploration that the Integrated Program Plan of the late 60s diverges most from the reality of the present. Currently, lunar exploration is confined to a several-hundred-kilometer radius around the lunar South Pole, and the IPP’s sketch of a nuclear-powered Mars mission in the 1980s is still far from being achieved. However, in both cases, planning and advocacy is already in place in several places to see these deficiencies rectified as the next stage of NASA’s exploration plans. The most basic of these suggestions is the further expansion of development at Armstrong, expanding the base and its permanent staff. Given that thanks to the LEO RLVs and the Gateway and TransOrbital depot and tug network, the total cost of a single seat on a 6-month rotation to Armstrong is lower than the cost sustained for Spacelab and Freedom using entirely expendable vehicle, this seems likely to occur as a minimum conservative estimate. More advanced concepts envision the development of a replacement for the nearly twenty-year-old Artemis lander design, incorporating reusability and depot-connection features. Such a design, combined with a Centaur or Pegasus used as a reusable “uncrasher” stage to transfer it to a lunar suborbital trajectory, could further decrease the cost of logistics and crew flights to Armstrong. A more ambitious proposal calls for a similar but larger vehicle using the same architecture for a reusable remote sortie architecture, landing a reusable “minibase” science station the size of the original Orion outpost for a few months at any point on the lunar surface before dusting off for orbit, returning to the wide-ranging exploration of Apollo and Artemis but with the vastly improved scientific capacity and decreased cost of Orion and Armstrong expeditions.

    However, while the further exploration of the Moon beckons to selenologists and members of O'Neill's Lunar Society, there is a strong drive to see the remaining destination in the Integrated Program Plan explored: Mars. Since the era of Apollo, Mars has been seen as the natural successor to the Moon for exploration and development, and the Red Planet’s attractions have only been boosted by television programs like Exodus and the ceaseless support of Robert Zubrin and his On to Mars! group. There is a strong case to be made--as it has been in both Zubrin and NASA-funded studies on mission concepts--that with NASA’s current Gateway network and experience with long duration spaceflight, Mars requires little more development than a renewed program of sortie “minibases’ with a reusable lander would on the Moon: one or two new vehicles, and a re-allocation or expansion of spending on human spaceflight. After all, a reusable Pegasus tug departing fully-fueled from an upgraded Gateway-2 station at EML-2 would be able to push a Saturn II-Heavy’s entire LEO payload to Earth escape, potentially enabling as much as 50 tons to be landed on the surface of Mars for less than half a billion dollars--a capability which makes expansive multi-launch Mars flights seem far more practical and less expensive than they were in the 1960s, when serious plans of Mars were almost certainly a step too far. Whether it is the Moon, Mars, or perhaps both which see the next phase of human exploration of space, the decreased costs and increased capacity of orbital infrastructure seem to suggest that the boundaries of Earth orbit shall not re-assert themselves any time soon.

    However, the current boom in commercial spaceflight means that NASA aren’t the only ones carving out their own plans and aspirations, nor are the additional plans limited to those of other traditional players like Europe, Russia, Japan, China, or India. As shown by the developing Korean program and recent announcements by Brazil and some of the Gulf States, the dropping price of launch and democratization of space access has enabled many nations to consider space exploration who previously were unable to do so. Moreover, this democratization of spaceflight isn’t limited just to governmental organizations, as shown by NEOSearch’s establishment of the first non-governmental space telescope for asteroid research and the success of semi-commercial tourist flights to Mir-II by Roscosmos. It remains to be seen if some of the ambitious schemes proposed for more extensive commercial spaceflight prove to be more than pipe dreams, from specialized “space hotels” for orbital tourists launched on second or third-generation reusables, asteroid resource extraction, privately-organized flights to the Moon or beyond, the development of lunar resources, or the construction of orbital power stations.

    Like the Integrated Program Plan of the past and like NASA’s own Moon and Mars aspirations of the present, these dreams are all within the realm of technical feasibility. The question of which, if any, may come to pass depends on political will and economic realities, as they have throughout the history of the space program. However, just as been true throughout history, there will no doubt be those who will find their inspiration with their eyes turned skyward.
     
    Eyes Turned Skyward: 10 Year Retrospective
  • Good evening, everyone! It’s been a while, hasn’t it? Ten years ago, Workable Goblin and I first started posting Eyes Turned Skyward here on AlternateHistory.com, a project which we’d been working on since the spring of 2011. First tip for other writers: pre-writing and editing before posting is a really useful model for technical timelines. Four parts, several hundred thousand words, and 4 years later, we posted the finale on September 11, 2015. Today, I wanted to take a moment to look back on why we wrote this timeline, what the situation in spaceflight was at the time, how that situation has changed, what we learned from Eyes, and what we might do the same or different if we were starting it from scratch today. However, before all of that, I wanted to start by saying thank you--thank you to the people who worked with us as we wrote this project, from our art team like Michel Van and Nixonshead, to some of our subject matter consultants like Brainbin, and also to the readers and commentators who’ve shared insights, thoughts on the timeline, made KSP mods and LEGO models, and so much more. I don’t know that I can count how many people I’ve gotten to know because of Eyes and how much time I’ve spent with friends I made through it. It was an amazing experience and a great project. We learned a lot about space, but also a lot about writing.

    What drove Eyes? Why did we decide to write a no-Shuttle timeline? I think we can agree, with hindsight, that 2011 represented a nadir in spaceflight. The cancellation of Constellation, while perhaps justifiable, at the time felt strikingly personal for me. Combined with the nearly simultaneous end of Shuttle, it felt like everything was coming down without much hope of something coming back. Although Workable Goblin had fewer personal problems with Constellation’s cancellation, he had a long interest in alternate Apollos and also felt that space had reached a dead end in the late 2000s, with human spaceflight seemingly on the verge of death. It felt like looking back to eliminate Shuttle was fixing the mistakes of the past. Basically no one in 2011 was seriously discussing reuse in public; not even SpaceX, which at the time was still talking about parachutes at most for Falcon 9. Reading Eyes and the author’s notes on posts, you can see how spaceflight has recovered and changed over the last decade--the rise of SpaceX, Dragon, Cygnus, recovery & reuse, the small launch market, the rebirth of LEO mega-constellations, and probes going to new destinations, proving each to be fascinating worlds to explore. In retrospect, though, we underestimated a bit how good the Shuttle could be. In writing Eyes and researching other projects since, it’s become easier to appreciate the benefits of the large crews Shuttle carried, and the economies it was capable of at higher flight rates when it achieved them. Today, we are once again looking at a bold new vehicle, SpaceX’s Starship, as well as other vehicles aiming for partial and full reusability, and wondering what the future of spaceflight may hold. I hope dearly it doesn’t blow up and fail in a way that causes a similar knee-jerk reaction in hindsight.

    At the time when we wrote Eyes, the state of the art for researching space timelines on AH.com was often, at most, Astronautix, typically without much in the way of calculations of capabilities or checks of original NTRS studies. At the time one of our major advantages was Workable Goblin’s PhD-student-honed abilities to dive into NTRS for studies on probes, stations, and more, and my college-engineering-student willingness to tinker with spreadsheets to try and game out modifications of paper rocket studies to line up capabilities. In the last decade, the availability of that material through David Portree’s work, False Steps, KSP mod research, and other mechanisms has exploded, bringing some of those possibilities more into the light.

    One notable result of that is highlighting perhaps one of the biggest weaknesses of the timeline: Saturn IC and Saturn Multibody. The flexibility of the system in the timeline was a big tool in the missions we called for, and some of the benefits of clusterable heavy lifters which have been realized in the real world. Still, in retrospect there are more than a few points of dissatisfaction with the rockets we laid out. The benefits of switching to the F-1 for Saturn IC, in planning, was that it made possible a 70-ton capable heavy lifter with only 3 engines, free from “N-1 syndrome”. On the other hand, as far as I’ve ever seen since, no historical sources ever really considered single-engine F-1-based core vehicles or liquid boosters, looking at a minimum at 6.6m twin-engine F-1 LRBs like Saturn MLV and the more modern Jarvis and Pyrios systems. Similarly, in the years since, the tremendous success of the Merlin and Falcon families shows perhaps N-1 Syndrome need not strike if you have better quality control and testing. Ironically, an early version of the thing which turned into Multibody in our planning would have had shared engines between an Atlas-sized, Saturn I-sized, and a roughly 70-ton LV, much like the historical H-1, so this was something we at least considered. Real aerospace engineers clearly saw a value in that which we overlooked in the drive for a single engine and saving the F-1. Additionally, while switching from the S-IB cluster tank to a monolithic 6.6m tank set might save weight, applying the Saturn V’s S-IC structural fraction to a stage a quarter the size was possibly unjustified, yielding a vehicle that performed just a bit too well. Wrapping it in SOFI orange on only the kerolox lower stage in one of the first early images of the Multibody family probably caps it off, as does never having a full answer for what roll control might have been like--turbopump tap off, secondary engines, or what--much to the frustration of KSP players replicating the vehicles.

    Likely, the changes from Saturn IB to Saturn IC in timeline wouldn’t have actually been worth it, and an interim Saturn IB should have flown, or Multibody-style flexibility and performance somehow introduced and justified earlier, probably with Altas V-style copperish anodization or a colored paint/coating. Maybe SOFI on the upper stage, but possibly in the white color used on early Shuttle or Saturn V’s S-II. A Titan alternative of some sort also gets talked about a lot in Shuttle decision documents, though it’s arguable that it would have had some serious capability limits in diameter of fairings and payload mass, not to mention that the costs the Air Force reported may have been somewhat manipulated to look cheaper through spares production for the missile fleet and such. There was a large-diameter Titan we considered for a while, but we ultimately preferred Multibody’s capability and kerolox first stage instead of hypergols. The issue, of course, is that untangling changes like these would require a major re-write of the timeline as it exists to fix, changing many of the details we think hold up. As we’ve seen with some authors lately, sometimes going back to long-completed projects and trying to edit them to fit changing times can collapse the whole endeavour. Thus, while we have points we’re dissatisfied with, I’m inclined to say the work stands as it was. It’s better to leave Saturn IC and Multibody as it is, and try to use what we’ve learned since for other projects instead of trying to re-write alternate history. In the eyes of many people just finding the timeline through Nixonshead’s images, KSP mods, or other methods, Multibody is Eyes Turned Skyward as written. If you’re one of them...welcome.

    Another issue in a similar vein was the success of Europa. This, we have to admit, was also very goal driven: keeping Britain as a major contributing member to ESA. But while Europa’s success was possible it wasn’t really likely, and this led to the development of a number of plot lines that were also unlikely and didn’t add very much to the timeline, like Lockheed’s success with the L-1011 and L-1012. This leads into another problem, which was the number of plot threads that ended up going nowhere or were tangential to the main thrust of the timeline. In some cases these could be somewhat justified--it was apparent pretty early on that we were going to have to tackle politics somehow just because of how the 2000 election hinged on Florida, one of the states most affected by the changes we had made. Overall, though, we probably spent too much time on subjects disconnected from the central space focus. I still recall going to a movie with my family the day we posted the politics update where Ann Richards became president...and got out an hour or two later to find a multi-post (multi-page?) argument about if she was a murderer for some case law around the death penalty in Texas while she was governor IOTL. Some of my later timelines have shied away from directly addressing or butterflying politics for reasons like these. While it can be interesting, it can stray from the core interest of playing with alternate rockets. We also abandoned certain interesting plot lines that probably deserved to get more focus than they actually did. For example, we spent a considerable amount of time in Part II discussing the rise of space advocacy organizations and then...did very little with them through the rest of the timeline, even though they should probably have had some impact on later decisions and definitely would have been strongly affected by the Artemis and later lunar exploration programs. But again, re-writing the timeline to fix these issues would take a large amount of work and risks ruining the appeal for both our readers and ourselves.

    The ideas that drove me in particular to write Eyes--wanting to imagine a better future for spaceflight--haven’t gone away, though. Similar drives have been behind most of the timelines I’ve written since, very much in an ongoing dialog with what we wrote in Eyes and in what I’ve learned since through research and watching developments in reality. Over the time we were Eyes, recovery & reuse of rockets emerged back onto the scene and have become something of a new default for LV proposals, which spurred me to look back through some of the history of Shuttle and reusable vehicle concepts. When you look at data like the history of the Shuttle program cost and particularly the 1994 Space Shuttle Zero-Base Cost study, it’s difficult to not feel a bit that what was more important about Shuttle being good or bad was not if it was reusable, but if it could truly fly 10-20 times a year. If Multibody really existed, it would likely be cheaper more thanks to flying something like ten times a year through the timeline than much about having only a single engine or only a single tank set on the first stage. This was an idea explored more fully in Right Side Up, looking at what a Shuttle program might have needed to routinely and safely that kind of flight rate, and then in Boldly Going this year trying to imagine what more might have been done starting with the Shuttle we got, with a station to drive flight rate like the 1998 peak earlier in the program’s history. I think my current largest unwritten idea, “Fires of Mercury,” would explore some of this with a radically overhauled early 1970s, in dialog with the assumption from Eyes of a convergent drawdown in Apollo, but also the benefits of flight frequency and continuing some use of Apollo hardware...but in a reusable way. The application of the moon in driving this flight rate, and as a natural place to return in exploration is now potentially being borne out in reality, as in Eyes and so many other of my timelines since. Even Dawn of the Dragon got in on the act, being the first of my timelines to feature multiple-launch EOR-LOR lunar flights using Shuttle-C as we’d used Saturn Multibody H03 in Eyes and also in a deliberate attempt to stay away from serious political discussion.

    In addition to concepts I continue to mine in other works to this day, Eyes also had some structural elements applied to other timelines. Some of these are lessons learned, like having more of a defined goal of when (if not necessarily where) to end a timeline from the beginning, and of defining what details were within the scope and detail of the timeline and which weren’t. While we enjoyed making Eyes’ culture posts and political butterflies, they also took up a lot of focus, and most of my later works have been much more tightly focused as a result. (Indeed, in editing a draft of this TimothyC reminded me that the first draft of Boldly Going eschewed discussing politics entirely.) Another lesson we learned was the benefit of pre-writing. We generally had a Part of Eyes about half written before going to post to enable a weekly schedule. On at least one occasion, this wasn’t enough, and edits & schoolwork bogged us down, leading to a hiatus mid-part to enable rebuilding a longer buffer. The buffer not only protected the schedule, it gave time for some of the “gaming out” and edits which I think helped build Eyes’ technical verisimilitude. That gaming-out of technical decisions, researching historical alternatives considered, brainstorming how these alternatives might change in the timeline, and then casting ourselves as program leaders making decisions year by year or even month by month in selecting launch order, landing sites, program priorities, and political objectives I think helped Eyes feel more real and grounded our technical justifications...even if some of the stuff we were justifying wasn’t always fully baked. That’s a method I’ve tried to take to my later works. I think it’s notable the only timeline I gave up on before finishing, Kistling a Different Tune, is one where I deliberately tried not pre-writing and didn’t game things out as much in advance of writing sections. Twists like the continued survival of Apollo, the approval of the Orion “Soonbase,” and the flight history of the Grumman X-40 in Eyes came from this process, and helped make the story more than the sum of its outline elements by linking across programs and nations. Letting the gaming-out drive the writing of the story outline, bounding what we could do with internally plausible limits, was I think a major element to making Eyes feel real, and I’ve tried to use it as much as possible in later timelines. Certainly the biggest impact of that was Boldly Going, where gaming out an April Fools joke timeline from 2020 intended to be five to seven thousand words inflated it to closer to seventy thousand words.

    Or course, the words of the timeline and the technical research behind its creation was only part of the timeline’s success. The art team we were able to work with was a major contribution to the spread of the timeline, I think. Every image from Eyes shared with a link to the thread was an advertisement for what was contained within, and there’s a reason I think Multibody, the surviving Apollo blocks, and the Space Station Freedom we imagined are some of the more public-facing elements of the timeline: they’re the easiest to show a quick link to promote. It invites explanation, questions, and I hope, reading and enjoying the timeline. Workable Goblin and I very much appreciate the effort of Nixonshead, Michel, and others who contributed to illustrating the world we created and helping to fill in details beyond the early Autodesk Inventor models I started the timeline with. I’ve tried to learn the lesson of the value of art in illustrating an AH premise in an enticing way for my other timelines, and consider myself fortunate to have continued to get to work with great art teams on other projects.

    Sometimes, the illustrations of the timeline’s events, like Apollo’s continuation, Space Station Freedom, the lunar return, and Multibody itself seem to be the most lasting impacts from Eyes more than the words we wrote. However, I’m proud of several things in Eyes that didn’t make it fully into the art, and which I’d suggest people look for if this drives them to re-read, or read in full for the first time! Workable Goblin dug up fantastic data about alternate probe missions, and took us on journeys to many new places, from the cryovolcanoes of Pluto before New Horizons helped prove they existed, to dusty little Phobos (still unexplored in the real world), to the scorched surface of Venus. Each was revealed in his writing to, as we have found in reality since, its own wonderful new world to explore, no matter how apparently barren they may appear at the start. He also managed to fully immerse himself in that school of acronyms that produced OTL’s GRAIL, and still cackles whenever he remembers GRIMALDI. I enjoy Kirchoff (an ion propelled multi-target probe we included with a name inspired both by a historical figure related to electricity but also derived from the real study name Comet Rendezvous/Comet Flyby: CR/CF). I’m pleased with bringing some human elements to our stories of the return to the moon, with some of the characters we introduced, and the scenes we set in Part III and Part IV. I’m also pleased with how we let the evolution of the story drive some decisions. Originally, Apollo was supposed to be replaced much earlier in the timeline, but it kept not becoming our NASA’s priority in a way which felt very real. Its final replacements then ended up making a nice bookend to the timeline with its return to the moon. The plans we developed for the evolution of the lunar return from week-long stays into bases consolidated at a specific site still feels like a solid plan. I’ll be interested to see with Artemis how much of any of this holds up. No Earth-Moon L2 staging, but an NRHO Gateway station, for one, and perhaps propellant transfer and reuse intended from the start instead of introduced in operations partway through? We shall see. It’s been an exciting decade, and there’s yet more to come. We are walking the earth with our Eyes Turned Skyward, hoping to look up and see a past that never was, or perhaps a future yet to be.
     
    Top