Space

Abadzis, Nick. Laika. New York: First Second, 2007. ISBN 978-1-59643-101-0.
The first living creature to orbit the Earth (apart, perhaps, from bacterial stowaways aboard Sputnik 1) was a tough, even-tempered, former stray dog from the streets of Moscow, named Kudryavka (Little Curly), who was renamed Laika (Barker) shortly before being sent on a one-way mission largely motivated by propaganda concerns and with only the most rudimentary biomedical monitoring in a slapdash capsule thrown together in less than a month.

This comic book (or graphic novel, if you prefer) tells the story through parallel narratives of the lives of Sergei Korolev, a former inmate of Stalin's gulag in Siberia who rose to be Chief Designer of the Soviet space program, and Kudryavka, a female part-Samoyed stray who was captured and consigned to the animal research section of the Soviet Institute of Aviation Medicine (IMBP). While obviously part of the story is fictionalised, for example Kudryavka's origin and life on the street, those parts of the narrative which are recorded in history are presented with scrupulous attention to detail. The author goes so far as to show the Moon in the correct phase in events whose dates are known precisely (although he does admit frankly to playing fast and loose with the time of moonrise and moonset for dramatic effect). This is a story of survival, destiny, ambition, love, trust, betrayal, empathy, cruelty, and politics, for which the graphic format works superbly—often telling the story entirely without words. For decades Soviet propaganda spread deception and confusion about Laika's fate. It was only in 2002 that Russian sources became available which revealed what actually happened, and the account here presents the contemporary consensus based upon that information.

March 2008 Permalink

Aldrin, Buzz. Magnificent Desolation. London: Bloomsbury, 2009. ISBN 978-1-4088-0416-2.
What do you do with the rest of your life when you were one of the first two humans to land on the Moon before you celebrated your fortieth birthday? This relentlessly candid autobiography answers that question for Buzz Aldrin (please don't write to chastise me for misstating his name: while born as Edwin Eugene Aldrin, Jr., he legally changed his name to Buzz Aldrin in 1979). Life after the Moon was not easy for Aldrin. While NASA trained their astronauts for every imaginable in-flight contingency, they prepared them in no way for their celebrity after the mission was accomplished, and detail-oriented engineers were suddenly thrust into the public sphere, sent as goodwill ambassadors around the world with little or no concern for the effects upon their careers or family lives.

All of this was not easy for Aldrin, and in this book he chronicles his marriages (3), divorces (2), battles against depression and alcoholism, search for a post-Apollo career, which included commanding the U.S. Air Force test pilot school at Edwards Air Force Base, writing novels, serving as a corporate board member, and selling Cadillacs. In the latter part of the book he describes his recent efforts to promote space tourism, develop affordable private sector access to space, and design an architecture which will permit exploration and exploitation of the resources of the Moon, Mars and beyond with budgets well below those of the Apollo era.

This book did not work for me. Buzz Aldrin has lived an extraordinary life: he developed the techniques for orbital rendezvous used to this day in space missions, pioneered underwater neutral buoyancy training for spacewalks then performed the first completely successful extra-vehicular activity on Gemini 12, demonstrating that astronauts can do useful work in the void, and was the second man to set foot on the Moon. But all of this is completely covered in the first three chapters, and then we have 19 more chapters describing his life after the Moon. While I'm sure it's fascinating if you've lived though it yourself, it isn't necessarily all that interesting to other people. Aldrin comes across as, and admits to being, self-centred, and this is much in evidence here. His adventures, ups, downs, triumphs, and disappointments in the post-Apollo era are those that many experience in their own lives, and I don't find them compelling to read just because the author landed on the Moon forty years ago.

Buzz Aldrin is not just an American hero, but a hero of the human species: he was there when the first naked apes reached out and set foot upon another celestial body (hear what he heard in his headphones during the landing). His life after that epochal event has been a life well-lived, and his efforts to open the high frontier to ordinary citizens are to be commended. This book is his recapitulation of his life so far, but I must confess I found the post-Apollo narrative tedious. But then, they wouldn't call him Buzz if there wasn't a buzz there! Buzz is 80 years old and envisions living another 20 or so. Works for me: I'm around 60, so that gives me 40 or so to work with. Given any remotely sane space policy, Buzz could be the first man to set foot on Mars in the next 15 years, and Lois could be the first woman. Maybe I and the love of my life will be among the crew to deliver them their supplies and the essential weasels for their planetary colonisation project.

A U.S. edition is available.

January 2011 Permalink

Aldrin, Buzz with Leonard David. Mission to Mars. Washington, National Geographic Society, 2013. ISBN 978-1-4262-1017-4.
As Buzz Aldrin (please don't write to chastise me for misstating his name: while born as Edwin Eugene Aldrin, Jr., he legally changed his name to Buzz Aldrin in 1988) notes, while Neil Armstrong may have been the first human to step onto the Moon, he was the first alien from another world to board a spacecraft bound for Earth (but how can he be sure?). After those epochal days in July of 1969, Aldrin, more than any other person who went to the Moon, has worked energetically to promote space exploration and settlement, developing innovative mission architectures to expand the human presence into the solar system. This work continues his intellectual contributions to human space flight which began with helping to develop the techniques of orbital rendezvous still employed today and pioneering neutral-buoyancy training for extra-vehicular activity, which enabled him to perform the first completely successful demonstration of work in orbit on Gemini XII.

In this book Aldrin presents his “Unified Space Vision” for the next steps beyond the home planet. He notes that what we know about the Moon today is very different from the little we mostly guessed when he set foot upon that world. Today it appears that the lunar polar regions may have abundant resources of water which provide not only a source of oxygen for lunar settlers, but electrolysed by abundant solar power, a source of rocket fuel for operations beyond the Earth. Other lunar resources may allow the fabrication of solar panels from in situ materials, reducing the mass which must be launched from the Earth. Aldrin “cyclers” will permit transfers between the Earth and Moon and the Earth and Mars with little expenditure of propellant.

Aldrin argues that space, from low Earth orbit to the vicinity of the Moon, be opened up to explorers, settlers, and entrepreneurs from all countries, private and governmental, to discover what works and what doesn't, and which activities make economic sense. To go beyond, however, he argues that the U.S. should take the lead, establishing a “United Strategic Space Enterprise” with the goal of establishing a permanent human settlement on Mars by 2035. He writes, “around 2020, every selected astronaut should consign to living out his or her life on the surface of Mars.”

And there's where it all falls apart for me. It seems to me the key question that is neither asked nor answered when discussing the establishment of a human settlement on Mars can be expressed in one word: “why?” Yes, I believe that long-term survival of humans and their descendants depends upon not keeping everything in one planetary basket, and I think there is tremendously interesting science to be done on Mars, which may inform us about the origin of life and its dissemination among celestial bodies, the cycle of climate on planets and the influence of the Sun, and many other fascinating subjects. It makes sense to have a number of permanent bases on Mars to study these things, just as the U.S. and other countries have maintained permanent bases in Antarctica for more than fifty years. But I no longer believe that the expansion of the human presence in the solar system is best accomplished by painfully clawing our way out of one deep gravity well only to make a long voyage and then make an extremely perilous descent into another one (the Martian atmosphere is thick enough you have to worry about entry heating, but not thick enough to help in braking to landing speed). Once you're on Mars, you only have solar power half the time, just as on Earth, and you have an atmosphere which is useless to breathe.

Even though few people take it seriously any more, Gerard K. O'Neill's vision of space settlements in The High Frontier (May 2013) makes far more sense to me. Despite Aldrin's enthusiasm for private space ventures, it seems to me that his vision for the exploration and settlement of Mars will be, for at least the first decades, the kind of elitist venture performed by civil servants that the Apollo Moon landings were. In this book he envisions no economic activity on Mars which would justify the cost of supporting an expanding human presence there. Now, wealthy societies may well fund a few bases, just as they do in the Antarctic, but that will never reach what O'Neill calls the point of “ignition”—where the settlement pays for itself and can fund its own expansion by generating economic value sufficient to import its needs and additional settlers. O'Neill works out in great detail how space settlements in cislunar space can do this, and I believe his economic case, first made in the 1970s, has not only never been refuted but is even more persuasive today.

Few people have thought as long and hard about what it takes to make our species a spacefaring civilisation as Buzz Aldrin, nor worked so assiduously over decades to achieve that goal. This is a concise summation of his view for where we should go from here. I disagree with much of his strategy, but hey, when it comes to extraterrestrial bodies, he's been there and I haven't. This is a slim book (just 272 pages in the hardback edition), and the last 20% is a time line of U.S. space policies by presidential administrations, including lengthy abstracts of speeches, quoted from space.com.

May 2013 Permalink

Bean, Alan and Andrew Chaikin. Apollo. Shelton, CT: The Greenwich Workshop, 1998. ISBN 978-0-86713-050-8.
On November 19th, 1969, Alan Bean became the fourth man to walk on the Moon, joining Apollo 12 commander Pete Conrad on the surface of Oceanus Procellarum. He was the first person to land on the Moon on his very first space flight. He later commanded the Skylab 3 mission in 1973, spending more than 59 days in orbit.

Astronauts have had a wide variety of second careers after retiring from NASA: executives, professors, politicians, and many others. Among the Apollo astronauts, only Alan Bean set out, after leaving NASA in 1981, to become a professional artist, an endeavour at which he has succeeded, both artistically and commercially. This large format coffee table book collects many of his paintings completed before its publication in 1998, with descriptions by the artist of the subject material of each and, in many cases, what he was trying to achieve artistically. The companion text by space writer Andrew Chaikin (A Man on the Moon) provides an overview of Bean's career and the Apollo program.

Bean's art combines scrupulous attention to technical detail (for example, the precise appearance of items reflected in the curved visor of spacesuit helmets) with impressionistic brushwork and use of colour, intended to convey how the lunar scenes felt, as opposed to the drab, near monochrome appearance of the actual surface. This works for some people, while others find it grating—I like it very much. Visit the Alan Bean Gallery and make up your own mind.

This book is out of print, but used copies are available. (While mint editions can be pricey, non-collector copies for readers just interested in the content are generally available at modest cost).

October 2008 Permalink

Benford, James and Gregory Benford, eds. Starship Century. Reno, NV: Lucky Bat Books, 2013. ISBN 978-1-939051-29-5.
“Is this the century when we begin to build starships?” So begins the book, produced in conjunction with the Starship Century Symposium held in May of 2013 at the University of California San Diego. Now, in a sense, we built and launched starships in the last century. Indeed, at this writing, eight objects launched from Earth are on interstellar trajectories. These are the two Pioneer spacecraft, the two Voyagers, the New Horizons Pluto flyby spacecraft, and its inert upper stage and two spin-down masses. But these objects are not aimed at any particular stars; they're simply flying outward from the solar system following whatever trajectory they were on when they completed their missions, and even if they were aimed at the nearest stars, it would take them tens of thousands of years to get there, by which time their radioactive power sources would be long exhausted and they would be inert space junk.

As long as they are built and launched by beings like humans (all bets are off should we pass the baton to immortal machines), starships or interstellar probes will probably need to complete their mission within the time scale of a human lifetime to be interesting. One can imagine multi-generation colony ships (and they are discussed here), but such ships are unlikely to be launched without confidence the destination is habitable, which can only be obtained by direct investigation by robotic probes launched previously. The closest star is around 4.3 light years from Earth. This is a daunting distance. To cross it in a human-scale time (say, within the career of a research scientist), you'd need to accelerate your probe to something on the order of 1/10 the speed of light. At this speed, each kilogram of the probe would have a kinetic energy of around 100 kilotons of TNT. A colony ship with a dry mass of 1,000 tonnes would, travelling at a tenth of the speed of light, have kinetic energy which, at a cost of USD 0.10 per kilowatt-hour, would be worth USD 12.5 trillion, which is impressive even by U.S. budget deficit standards. But you can't transmit energy to a spacecraft with 100% efficiency (the power cord is a killer!), and so the cost of a realistic mission might be ten times this.

Is it then, silly, to talk about starships? Well, not so fast. Ever since the Enlightenment, the GDP per capita has been rising rapidly. When I was a kid, millionaires were exotic creatures, while today people who bought houses in coastal California in the 1970s are all millionaires. Now it's billionaires who are the movers and shakers, and some of them are using their wealth to try to reduce the cost of access to space. (Yes, currency depreciation has accounted for a substantial part of the millionaire to billionaire transition, but the scope of what one can accomplish with a billion dollar grubstake today is still much greater than with a million dollars fifty years ago.) If this growth continues, might it not be possible that before this century is out there will be trillionaires who, perhaps in a consortium, have the ambition to expand the human presence to other stars?

This book collects contributions from those who have thought in great detail about the challenges of travel to the stars, both in nuts and bolts hardware and economic calculations and in science fictional explorations of what it will mean for the individuals involved and the societies which attempt that giant leap. There are any number of “Aha!” moments here. Freeman Dyson points out that the void between the stars is not as empty as many imagine it to be, but filled with Oort cloud objects which may extend so far as to overlap the clouds of neighbouring stars. Dyson imagines engineered organisms which could render these bodies habitable to (perhaps engineered) humans, which would expand toward the stars much like the Polynesians in the Pacific: from island to island, with a population which would dwarf both in numbers and productivity that of the inner system rock where they originated.

We will not go to the stars with rockets like we use today. The most rudimentary working of the numbers shows how absurd that would be. And yet nuclear thermal rockets, a technology developed and tested in the 1960s and 1970s, are more than adequate to develop a solar system wide economy which could support interstellar missions. Many different approaches to building starships are explored here: some defy the constraints of the rocket equation by keeping the power source in the solar system, as in “sailships” driven by laser or microwave radiation. A chapter explores “exotic propulsion”, beyond our present understanding of physics, which might change the game. (And before you dismiss such speculations, recall that according to the consensus model of cosmology, around 95% of the universe is made up of “dark matter” and “dark energy” whose nature is entirely unknown. Might it be possible that a vacuum propeller could be discovered which works against these pervasive media just as a submarine's propeller acts upon the ocean?)

Leavening the technical articles are science fiction stories exploring the transition from a planetary species to the stars. Science fiction provides the dreams which are then turned into equations and eventually hardware, and it has a place at this table. Indeed, many of the scientists who spoke at the conference and authored chapters in this book also write science fiction. We are far from being able to build starships or even interstellar probes but, being human, we're always looking beyond the horizon and not just imagining what's there but figuring out how we'll go and see it for ourselves. To date, humans haven't even learned how to live in space: our space stations are about camping in space, with extensive support from the Earth. We have no idea what it takes to create a self-sustaining closed ecosystem (consider that around 90% of the cells in your body are not human but rather symbiotic microbes: wouldn't you just hate it to be half way to Alpha Centauri and discover you'd left some single-celled critter behind?). If somebody waved a magic wand and handed us a propulsion module that could take us to the nearest stars within a human lifetime, there are many things we'd still need to know in order to expect to survive the journey and establish ourselves when we arrived. And, humans being humans, we'd go anyway, regardless. Gotta love this species!

This is an excellent survey of current thinking about interstellar missions. If you're interested in this subject, be sure to view the complete video archive of the conference, which includes some presentations which do not figure in this volume, including the magnificent galaxy garden.

November 2013 Permalink

Miller, Ron and Fredrick C. Durant III. The Art of Chesley Bonestell. London: Paper Tiger, 2001. ISBN 978-1-85585-884-8.
If you're interested in astronomy and space, you're almost certainly familiar with the space art of Chesley Bonestell, who essentially created the genre of realistic depictions of extraterrestrial scenes. But did you know that Bonestell also:

  • Was a licensed architect in the State of California, who contributed to the design of a number of buildings erected in Northern California in the aftermath of the 1906 earthquake?
  • Chose the site for the 1915 Panama-Pacific International Exposition (of which the San Francisco Palace of Fine Arts remains today)?
  • Laid out the Seventeen Mile Drive in Pebble Beach on the Monterey Peninsula?
  • Did detailed design of the ornamentation of the towers of the Golden Gate Bridge, and illustrated pamphlets explaining the engineering of the bridge?
  • Worked for years in Hollywood doing matte paintings for films including Citizen Kane?
  • Not only did the matte paintings, but designed the buildings of Howard Roark for the film version of The Fountainhead?
  • Painted the Spanish missions of California as they would have appeared in their heyday?

Although Bonestell always considered himself an illustrator, not an artist, and for much of his career took no particular care to preserve the originals of his work, here was a polymath with a paintbrush who brought genius as well as precision to every subject he rendered. He was, like his collaborator on Destination Moon, Robert A. Heinlein (the two admired each other's talents, but Bonestell thought Heinlein somewhat of a nut in his political views; their relationship got off to a rocky start when Bonestell visited Heinlein's self-designed dream house and pronounced his architectural judgement that it looked like a gas station), a businessman first—he would take the job that paid best and quickest, and produced a large volume of commercial art to order, all with the attention to detail of his more artistically ambitious creations.

While Bonestell was modest about his artistic pretensions, he had no shortage of self-esteem: in 1974 he painted a proposed redesign of the facade of St. Peter's Basilica better in keeping with his interpretation of Michelangelo's original intent and arranged to have it sent to the Pope who responded, in essence, “Thanks, but no thanks”.

This resplendent large-format coffee table book tells the story of Bonestell's long and extraordinarily creative career in both text and hundreds of full-colour illustrations of his work. To open this book to almost any page is to see worlds unknown at the time, rendered through the eye of an artist whose mind transported him there and sparked the dream of exploration in the generations which expanded the human presence and quest to explore beyond the home planet.

This book is out of print and used copies command a frightful premium; I bought this book when it was for sale at the cover price and didn't get around to reading all the text for seven years, hence its tardy appearance here.

November 2008 Permalink

Byers, Bruce K. Destination Moon. Washington: National Aeronautics and Space Administration, 1977. NASA TM X-3487.
In the mid 1960s, the U.S. Apollo lunar landing program was at the peak of its budget commitment and technical development. The mission mode had already been chosen and development of the flight hardware was well underway, along with the ground infrastructure required to test and launch it and the global network required to track missions in flight. One nettlesome problem remained. The design of the lunar module made assumptions about the properties of the lunar surface upon which it would alight. If the landing zone had boulders which were too large, craters sufficiently deep and common that the landing legs could not avoid, or slopes too steep to avoid an upset on landing or tipping over afterward, lunar landing missions would all be aborted by the crew when they reached decision height, judging there was no place they could set down safely. Even if all the crews returned safely without having landed, this would be an ignominious end to the ambitions of Project Apollo.

What was needed in order to identify safe landing zones was high-resolution imagery of the Moon. The most capable Earth-based telescopes, operating through Earth's turbulent and often murky atmosphere, produced images which resolved objects at best a hundred times larger that those which could upset a lunar landing mission. What was required was a large area, high resolution mapping of the Moon and survey of potential landing zones, which could only be done, given the technology of the 1960s, by going there, taking pictures, and returning them to Earth. So was born the Lunar Orbiter program, which in 1966 and 1967 sent lightweight photographic reconnaissance satellites into lunar orbit, providing both the close-up imagery needed to select landing sites for the Apollo missions, but also mapping imagery which covered 99% of the near side of the Moon and 85% of the far side, In fact, Lunar Orbiter provided global imagery of the Moon far more complete than that which would be available for the Earth many years thereafter.

Accomplishing this goal with the technology of the 1960s was no small feat. Electronic imaging amounted to analogue television, which, at the altitude of a lunar orbit, wouldn't produce images any better than telescopes on Earth. The first spy satellites were struggling to return film from Earth orbit, and returning film from the Moon was completely impossible given the mass budget of the launchers available. After a fierce competition, NASA contracted with Boeing to build the Lunar Orbiter, designed to fit on NASA's workhorse Atlas-Agena launcher, which seriously constrained its mass. Boeing subcontracted with Kodak to build the imaging system and RCA for the communications hardware which would relay the images back to Earth and allow the spacecraft to be controlled from the ground.

The images were acquired by a process which may seem absurd to those accustomed to present-day digital technologies but which seemed miraculous in its day. In lunar orbit, the spacecraft would aim its cameras (it had two: a mapping camera which produced overlapping wide-angle views and a high-resolution camera that photographed clips of each frame with a resolution of about one metre) at the Moon and take a series of photos. Because the film used had a very low light sensitivity (ASA [now ISO] 1.6), on low-altitude imaging passes the film would have to be moved to compensate for the motion of the spacecraft to avoid blurring. (The low light sensitivity of the film was due to its very high spatial resolution, but also reduced its likelihood of being fogged by exposure to cosmic rays or energetic particles from solar flares.)

After being exposed, the film would subsequently be processed on-board by putting it in contact with a band containing developer and fixer, and then the resulting negative would be read back for transmission to Earth by scanning it with a moving point of light, measuring the transmission through the negative, and sending the measured intensity back as an analogue signal. At the receiving station, that signal would be used to modulate the intensity of a spot of light scanned across film which, when developed and assembled into images from strips, revealed the details of the Moon. The incoming analogue signal was recorded on tape to provide a backup for the film recording process, but nothing was done with the tapes at the time. More about this later….

Five Lunar Orbiter missions were launched, and although some experienced problems, all achieved their primary mission objectives. The first three missions provided all of the data required by Apollo, so the final two could be dedicated to mapping the Moon from near-polar orbits. After the completion of their primary imaging missions, Lunar Orbiters continued to measure the radiation and micrometeoroid environment near the Moon, and contributed to understanding the Moon's gravitational field, which would be important in planning later Apollo missions that would fly in very low orbits around the Moon. On August 23rd, 1966, the first Lunar Orbiter took one of the most iconic pictures of the 20th century: Earthrise from the Moon. The problems experienced by Lunar Orbiter missions and the improvisation by ground controllers to work around them set the pattern for subsequent NASA robotic missions, with their versatile, reconfigurable flight hardware and fine-grained control from the ground.

You might think the story of Lunar Orbiter a footnote to space exploration history which has scrolled off the screen with subsequent Apollo lunar landings and high-resolution lunar mapping by missions such as Clementine and the Lunar Reconnaissance Orbiter, but that fails to take into account the exploits of 21st century space data archaeologists. Recall that I said that all of the image data from Lunar Orbiter missions was recorded on analogue tapes. These tapes contained about 10 bits of dynamic range, as opposed to the 8 bits which were preserved by the optical recording process used in receiving the images during the missions. This, combined with contemporary image processing techniques, makes for breathtaking images recorded almost half a century ago, but never seen before. Here are a document and video which record the exploits of the Lunar Orbiter Image Recovery Project (LOIRP). Please visit the LOIRP Web site for more restored images and details of the process of restoration.

September 2014 Permalink

Cabbage, Michael and William Harwood. Comm Check…The Final Flight of Shuttle Columbia. New York: Free Press, 2004. ISBN 0-7432-6091-0.
This is an excellent account for the general reader of the Space Shuttle Columbia STS-107 accident and subsequent investigation. The authors are veteran space reporters: Cabbage for the Orlando Sentinel and Harwood for CBS News. If you've already read the Columbia Accident Investigation Board Report (note that supplementary volumes II through VI are now available), you won't learn anything new about the technical details of the accident and its engineering and organisational causes here, but there's interesting information about the dynamics of the investigation and the individuals involved which you won't find in the formal report. The NASA Implementation Plan for Return to Flight and Beyond mentioned on page 264 is available online.

October 2004 Permalink

Cadbury, Deborah. Space Race. London: Harper Perennial, 2005. ISBN 0-00-720994-0.
This is an utterly compelling history of the early years of the space race, told largely through the parallel lives of mirror-image principals Sergei Korolev (anonymous Chief Designer of the Soviet space program, and beforehand slave labourer in Stalin's Gulag) and Wernher von Braun, celebrity driving force behind the U.S. push into space, previously a Nazi party member, SS officer, and user of slave labour to construct his A-4/V-2 weapons. Drawing upon material not declassified by the United States until the 1980s and revealed after the collapse of the Soviet Union, the early years of these prime movers of space exploration are illuminated, along with how they were both exploited by and deftly manipulated their respective governments. I have never seen the story of the end-game between the British, Americans, and Soviets to spirit the V-2 hardware, technology, and team from Germany in the immediate post-surrender chaos told so well in a popular book. The extraordinary difficulties of trying to get things done in the Soviet command economy are also described superbly, and underline how inspired and indefatigable Korolev must have been to accomplish what he did.

Although the book covers the 1930s through the 1969 Moon landing, the main focus is on the competition between the U.S. and the Soviet Union between the end of World War II and the mid-1960s. Out of 345 pages of main text, the first 254 are devoted to the period ending with the flights of Yuri Gagarin and Alan Shepard in 1961. But then, that makes sense, given what we now know about the space race (and you'll know, if you don't already, after reading this book). Although nobody in the West knew at the time, the space race was really over when the U.S. made the massive financial commitment to Project Apollo and the Soviets failed to match it. Not only was Korolev compelled to work within budgets cut to half or less of his estimated requirements, the modest Soviet spending on space was divided among competing design bureaux whose chief designers engaged in divisive and counterproductive feuds. Korolev's N-1 Moon rocket used 30 first stage engines designed by a jet engine designer with modest experience with rockets because Korolev and supreme Soviet propulsion designer Valentin Glushko were not on speaking terms, and he was forced to test the whole grotesque lash-up for the first time in flight, as there wasn't the money for a ground test stand for the complete first stage. Unlike the “all-up” testing of the Apollo-Saturn program, where each individual component was exhaustively ground tested in isolation before being committed to flight, it didn't work. It wasn't just the Soviets who took risks in those wild and wooly days, however. When an apparent fuel leak threatened to delay the launch of Explorer-I, the U.S. reply to Sputnik, brass in the bunker asked for a volunteer “without any dependants” to go out and scope out the situation beneath the fully-fuelled rocket, possibly leaking toxic hydrazine (p. 175).

There are a number of factual goofs. I'm not sure the author fully understands orbital mechanics which is, granted, a pretty geeky topic, but one which matters when you're writing about space exploration. She writes that the Jupiter C re-entry experiment reached a velocity (p. 154) of 1600 mph (actually 16,000 mph), that Yuri Gararin's Vostok capsule orbited (p. 242) at 28,000 mph (actually 28,000 km/h), and that if Apollo 8's service module engine had failed to fire after arriving at the Moon (p. 325), the astronauts “would sail on forever, lost in space” (actually, they were on a “free return” trajectory, which would have taken them back to Earth even if the engine failed—the critical moment was actually when they fired the same engine to leave lunar orbit on Christmas Day 1968, which success caused James Lovell to radio after emerging from behind the Moon after the critical burn, “Please be informed, there is a Santa Claus”). Orbital attitude (the orientation of the craft) is confused with altitude (p. 267), and retro-rockets are described as “breaking rockets” (p. 183)—let's hope not! While these and other quibbles will irk space buffs, they shouldn't deter you from enjoying this excellent narrative.

A U.S. edition is now available. The author earlier worked on the production of a BBC docu-drama also titled Space Race, which is now available on DVD. Note, however, that this is a PAL DVD with a region code of 2, and will not play unless you have a compatible DVD player and television; I have not seen this programme.

October 2007 Permalink

Carpenter, [Malcolm] Scott and Kris Stoever. For Spacious Skies. New York: Harcourt, 2002. ISBN 0-15-100467-6.
This is the most detailed, candid, and well-documented astronaut memoir I've read (Collins' Carrying the Fire is a close second). Included is a pointed riposte to “the man malfunctioned” interpretation of Carpenter's MA-7 mission given in Chris Kraft's autobiography Flight (May 2001). Co-author Stoever is Carpenter's daughter.

June 2003 Permalink

Carroll, Michael. Living Among Giants. Cham, Switzerland: Springer International, 2015. ISBN 978-3-319-10673-1.
In school science classes, we were taught that the solar system, our home in the galaxy, is a collection of planets circling a star, along with assorted debris (asteroids, comets, and interplanetary dust). Rarely did we see a representation of either the planets or the solar system to scale, which would allow us to grasp just how different various parts of the solar system are from another. (For example, Jupiter is more massive than all the other planets and their moons combined: a proud Jovian would probably describe the solar system as the Sun, Jupiter, and other detritus.)

Looking more closely at the solar system, with the aid of what has been learned from spacecraft exploration in the last half century, results in a different picture. The solar system is composed of distinct neighbourhoods, each with its own characteristics. There are four inner “terrestrial” or rocky planets: Mercury, Venus, Earth, and Mars. These worlds huddle close to the Sun, bathing in its lambent rays. The main asteroid belt consists of worlds like Ceres, Vesta, and Pallas, all the way down to small rocks. Most orbit between Mars and Jupiter, and the feeble gravity of these bodies and their orbits makes it relatively easy to travel from one to another if you're patient.

Outside the asteroid belt is the domain of the giants, which are the subject of this book. There are two gas giants: Jupiter and Saturn, and two ice giants: Uranus and Neptune. Distances here are huge compared to the inner solar system, as are the worlds themselves. Sunlight is dim (at Saturn, just 1% of its intensity at Earth, at Neptune 1/900 that at Earth). The outer solar system is not just composed of the four giant planets: those planets have a retinue of 170 known moons (and doubtless many more yet to be discovered), which are a collection of worlds as diverse as anywhere else in the domain of the Sun: there are sulfur-spewing volcanos, subterranean oceans of salty water, geysers, lakes and rain of hydrocarbons, and some of the most spectacular terrain and geology known. Jupiter's moon Ganymede is larger than the planet Mercury, and appears to have a core of molten iron, like the Earth.

Beyond the giants is the Kuiper Belt, with Pluto its best known denizen. This belt is home to a multitude of icy worlds—statistical estimates are that there may be as many as 700 undiscovered worlds as large or larger than Pluto in the belt. Far more distant still, extending as far as two light-years from the Sun, is the Oort cloud, about which we know essentially nothing except what we glean from the occasional comet which, perturbed by a chance encounter or passing star, plunges into the inner solar system. With our present technology, objects in the Oort cloud are utterly impossible to detect, but based upon extrapolation from comets we've observed, it may contain trillions of objects larger than one kilometre.

When I was a child, the realm of the outer planets was shrouded in mystery. While Jupiter, Saturn, and Uranus can be glimpsed by the unaided eye (Uranus, just barely, under ideal conditions, if you know where to look), and Neptune can be spotted with a modest telescope, the myriad moons of these planets were just specks of light through the greatest of Earth-based telescopes. It was not until the era of space missions to these worlds, beginning with the fly-by probes Pioneer and Voyager, then the orbiters Galileo and Cassini, that the wonders of these worlds were revealed.

This book, by science writer and space artist Michael Carroll, is a tourist's and emigrant's guide to the outer solar system. Everything here is on an extravagant scale, and not always one hospitable to frail humans. Jupiter's magnetic field is 20,000 times stronger than that of Earth and traps radiation so intense that astronauts exploring its innermost large moon Io would succumb to a lethal dose of radiation in minutes. (One planetary scientist remarked, “You need to have a good supply of grad students when you go investigate Io.”) Several of the moons of the outer planets appear to have oceans of liquid water beneath their icy crust, kept liquid by tidal flexing as they orbit their planet and interact with other moons. Some of these oceans may contain more water than all of the Earth's oceans. Tidal flexing may create volcanic plumes which inject heat and minerals into these oceans. On Earth, volcanic vents on the ocean floor provide the energy and nutrients for a rich ecosystem of life which exists independent of the Sun's energy. On these moons—who knows? Perhaps some day we shall explore these oceans in our submarines and find out.

Saturn's moon Titan is an amazing world. It is larger than Mercury, and has an atmosphere 50% denser than the Earth's, made up mostly of nitrogen. It has rainfall, rivers, and lakes of methane and ethane, and at its mean temperature of 93.7°K, water ice is a rock as hard as granite. Unique among worlds in the solar system, you could venture outside your space ship on Titan without a space suit. You'd need to dress very warmly, to be sure, and wear an oxygen mask, but you could explore the shores, lakes, and dunes of Titan protected only against the cold. With the dense atmosphere and gravity just 85% of that of the Earth's Moon, you might be able to fly with suitable wings.

We have had just a glimpse of the moons of Uranus and Neptune as Voyager 2 sped through their systems on its way to the outer darkness. Further investigation will have to wait for orbiters to visit these planets, which probably will not happen for nearly two decades. What Voyager 2 saw was tantalising. On Uranus's moon Miranda, there are cliffs 14 km high. With the tiny gravity, imagine the extreme sports you could do there! Neptune's moon Triton appears to be a Kuiper Belt object captured into orbit around Neptune and, despite its cryogenic temperature, appears to be geologically active.

There is no evidence for life on any of these worlds. (Still, one wonders about those fish in the dark oceans.) If barren, “all these worlds are ours”, and in the fullness of time we shall explore, settle, and exploit them to our own ends. The outer solar system is just so much bigger and more grandiose than the inner. It's as if we've inhabited a small island for all of our history and, after making a treacherous ocean voyage, discovered an enormous empty continent just waiting for us. Perhaps in a few centuries residents of these remote worlds will look back toward the Sun, trying to spot that pale blue dot so close to it where their ancestors lived, and remark to their children, “Once, that's all there was.”

March 2015 Permalink

Chaikin, Andrew. John Glenn: America's Astronaut. Washington: Smithsonian Books, 2014. ISBN 978-1-58834-486-1.
This short book (around 126 pages print equivalent), available only for the Kindle as a “Kindle single” at a modest price, chronicles the life and space missions of the first American to orbit the Earth. John Glenn grew up in a small Ohio town, the son of a plumber, and matured during the first great depression. His course in life was set when, in 1929, his father took his eight year old son on a joy ride offered by a pilot at local airfield in a Waco biplane. After that, Glenn filled up his room with model airplanes, intently followed news of air racers and pioneers of exploration by air, and in 1938 attended the Cleveland Air Races. There seemed little hope of his achieving his dream of becoming an airman himself: pilot training was expensive, and his family, while making ends meet during the depression, couldn't afford such a luxury.

With the war in Europe underway and the U.S. beginning to rearm and prepare for possible hostilities, Glenn heard of a government program, the Civilian Pilot Training Program, which would pay for his flying lessons and give him college credit for taking them. He entered the program immediately and received his pilot's license in May 1942. By then, the world was a very different place. Glenn dropped out of college in his junior year and applied for the Army Air Corps. When they dawdled accepting him, he volunteered for the Navy, which immediately sent him to flight school. After completing advanced flight training, he transferred to the Marine Corps, which was seeking aviators.

Sent to the South Pacific theatre, he flew 59 combat missions, mostly in close air support of ground troops in which Marine pilots specialise. With the end of the war, he decided to make the Marines his career and rotated through a number of stateside posts. After the outbreak of the Korean War, he hoped to see action in the jet combat emerging there and in 1953 arrived in country, again flying close air support. But an exchange program with the Air Force finally allowed him to achieve his ambition of engaging in air to air combat at ten miles a minute. He completed 90 combat missions in Korea, and emerged as one of the Marine Corps' most distinguished pilots.

Glenn parlayed his combat record into a test pilot position, which allowed him to fly the newest and hottest aircraft of the Navy and Marines. When NASA went looking for pilots for its Mercury manned spaceflight program, Glenn was naturally near the top of the list, and was among the 110 military test pilots invited to the top secret briefing about the project. Despite not meeting all of the formal selection criteria (he lacked a college degree), he performed superbly in all of the harrowing tests to which candidates were subjected, made cut after cut, and was among the seven selected to be the first astronauts.

This book, with copious illustrations and two embedded videos, chronicles Glenn's career, his harrowing first flight into space, his 1998 return to space on Space Shuttle Discovery on STS-95, and his 24 year stint in the U.S. Senate. I found the picture of Glenn after his pioneering flight somewhat airbrushed. It is said that while in the Senate, “He was known as one of NASA's strongest supporters on Capitol Hill…”, and yet in fact, while not one of the rabid Democrats who tried to kill NASA like Walter Mondale, he did not speak out as an advocate for a more aggressive space program aimed at expanding the human presence in space. His return to space is presented as the result of his assiduously promoting the benefits of space research for gerontology rather than a political junket by a senator which would generate publicity for NASA at a time when many people had tuned out its routine missions. (And if there was so much to be learned by flying elderly people in space, why was it never done again?)

John Glenn was a quintessential product of the old, tough America. A hero in two wars, test pilot when that was one of the most risky of occupations, and first to ride the thin-skinned pressure-stabilised Atlas rocket into orbit, his place in history is assured. His subsequent career as a politician was not particularly distinguished: he initiated few pieces of significant legislation and never became a figure on the national stage. His campaign for the 1984 Democratic presidential nomination went nowhere, and he was implicated in the “Keating Five” scandal. John Glenn accomplished enough in the first forty-five years of his life to earn him a secure place in American history. This book does an excellent job of recounting those events and placing them in the context of the time. If it goes a bit too far in lionising his subsequent career, that's understandable: a biographer shouldn't always succumb to balance when dealing with a hero.

April 2014 Permalink

Chertok, Boris E. Rockets and People. Vol. 1. Washington: National Aeronautics and Space Administration, [1999] 2005. ISBN 978-1-4700-1463-6 NASA SP-2005-4110.
This is the first book of the author's monumental four-volume autobiographical history of the Soviet missile and space program. Boris Chertok was a survivor, living through the Bolshevik revolution, Stalin's purges of the 1930s, World War II, all of the postwar conflict between chief designers and their bureaux and rival politicians, and the collapse of the Soviet Union. Born in Poland in 1912, he died in 2011 in Moscow. After retiring from the RKK Energia organisation in 1992 at the age of 80, he wrote this work between 1994 and 1999. Originally published in Russian in 1999, this annotated English translation was prepared by the NASA History Office under the direction of Asif A. Siddiqi, author of Challenge to Apollo (April 2008), the definitive Western history of the Soviet space program.

Chertok saw it all, from the earliest Soviet experiments with rocketry in the 1930s, uncovering the secrets of the German V-2 amid the rubble of postwar Germany (he was the director of the Institute RABE, where German and Soviet specialists worked side by side laying the foundations of postwar Soviet rocketry), the glory days of Sputnik and Gagarin, the anguish of losing the Moon race, and the emergence of Soviet preeminence in long-duration space station operations.

The first volume covers Chertok's career up to the conclusion of his work in Germany in 1947. Unlike Challenge to Apollo, which is a scholarly institutional and technical history (and consequently rather dry reading), Chertok gives you a visceral sense of what it was like to be there: sometimes chilling, as in his descriptions of the 1930s where he matter-of-factly describes his supervisors and colleagues as having been shot or sent to Siberia just as an employee in the West would speak of somebody being transferred to another office, and occasionally funny, as when he recounts the story of the imperious Valentin Glushko showing up at his door in a car belching copious smoke. It turns out that Glushko had driven all the way with the handbrake on, and his subordinate hadn't dared mention it because Glushko didn't like to be distracted when at the wheel.

When the Soviets began to roll out their space spectaculars in the late 1950s and early '60s, some in the West attributed their success to the Soviets having gotten the “good German” rocket scientists while the West ended up with the second team. Chertok's memoir puts an end to such speculation. By the time the Americans and British vacated the V-2 production areas, they had packed up and shipped out hundreds of rail cars of V-2 missiles and components and captured von Braun and all of his senior staff, who delivered extensive technical documentation as part of their surrender. This left the Soviets with pretty slim pickings, and Chertok and his staff struggled to find components, documents, and specialists left behind. This put them at a substantial disadvantage compared to the U.S., but forced them to reverse-engineer German technology and train their own people in the disciplines of guided missilery rather than rely upon a German rocket team.

History owes a great debt to Boris Chertok not only for the achievements in his six decade career (for which he was awarded Hero of Socialist Labour, the Lenin Prize, the Order of Lenin [twice], and the USSR State Prize), but for living so long and undertaking to document the momentous events he experienced at the first epoch at which such a candid account was possible. Only after the fall of the Soviet Union could the events chronicled here be freely discussed, and the merits and shortcomings of the Soviet system in accomplishing large technological projects be weighed.

As with all NASA publications, the work is in the public domain, and an online PDF edition is available.

A Kindle edition is available which is perfectly readable but rather cheaply produced. Footnotes simply appear in the text in-line somewhere after the reference, set in small red type. Words are occasionally run together and capitalisation is missing on some proper nouns. The index references page numbers from the print edition which are not included in the Kindle version, and hence are completely useless. If you have a workable PDF application on your reading device, I'd go with the NASA PDF, which is not only better formatted but free.

The original Russian edition is available online.

May 2012 Permalink

Chertok, Boris E. Rockets and People. Vol. 2. Washington: National Aeronautics and Space Administration, [1999] 2006. ISBN 978-1-4700-1508-4 NASA SP-2006-4110.
This is the second book of the author's four-volume autobiographical history of the Soviet missile and space program. Boris Chertok was a survivor, living through the Bolshevik revolution, the Russian civil war, Stalin's purges of the 1930s, World War II, all of the postwar conflict between chief designers and their bureaux and rival politicians, and the collapse of the Soviet Union. Born in Poland in 1912, he died in 2011 in Moscow. After retiring from the RKK Energia organisation in 1992 at the age of 80, he wrote this work between 1994 and 1999. Originally published in Russian in 1999, this annotated English translation was prepared by the NASA History Office under the direction of Asif A. Siddiqi, author of Challenge to Apollo (April 2008), the definitive Western history of the Soviet space program.

Volume 2 of Chertok's chronicle begins with his return from Germany to the Soviet Union, where he discovers, to his dismay, that day-to-day life in the victorious workers' state is much harder than in the land of the defeated fascist enemy. He becomes part of the project, mandated by Stalin, to first launch captured German V-2 missiles and then produce an exact Soviet copy, designated the R-1. Chertok and his colleagues discover that making a copy of foreign technology may be more difficult than developing it from scratch—the V-2 used a multitude of steel and non-ferrous metal alloys, as well as numerous non-metallic components (seals, gaskets, insulation, etc.) which were not produced by Soviet industry. But without the experience of the German rocket team (which, by this time, was in the United States), there was no way to know whether the choice of a particular material was because its properties were essential to its function or simply because it was readily available in Germany. Thus, making an “exact copy” involved numerous difficult judgement calls where the designers had to weigh the risk of deviation from the German design against the cost of standing up a Soviet manufacturing capacity which might prove unnecessary.

After the difficult start which is the rule for missile projects, the Soviets managed to turn the R-1 into a reliable missile and, through patience and painstaking analysis of telemetry, solved a mystery which had baffled the Germans: why between 10% and 20% of V-2 warheads had detonated in a useless airburst high above the intended target. Chertok's instrumentation proved that the cause was aerodynamic heating during re-entry which caused the high explosive warhead to outgas, deform, and trigger the detonator.

As the Soviet missile program progresses, Chertok is a key player, participating in the follow-on R-2 project (essentially a Soviet Redstone—a V-2 derivative, but entirely of domestic design), the R-5 (an intermediate range ballistic missile eventually armed with nuclear warheads), and the R-7, the world's first intercontinental ballistic missile, which launched Sputnik, Gagarin, and whose derivatives remain in service today, providing the only crewed access to the International Space Station as of this writing.

Not only did the Soviet engineers have to build ever larger and more complicated hardware, they essentially had to invent the discipline of systems engineering all by themselves. While even in aviation it is often possible to test components in isolation and then integrate them into a vehicle, working out interface problems as they manifest themselves, in rocketry everything interacts, and when something goes wrong, you have only the telemetry and wreckage upon which to base your diagnosis. Consider: a rocket ascending may have natural frequencies in its tankage structure excited by vibration due to combustion instabilities in the engine. This can, in turn, cause propellant delivery to the engine to oscillate, which will cause pulses in thrust, which can cause further structural stress. These excursions may cause control actuators to be over-stressed and possibly fail. When all you have to go on is a ragged cloud in the sky, bits of metal raining down on the launch site, and some telemetry squiggles for a second or two before everything went pear shaped, it can be extraordinarily difficult to figure out what went wrong. And none of this can be tested on the ground. Only a complete systems approach can begin to cope with problems like this, and building that kind of organisation required a profound change in Soviet institutions, which had previously been built around imperial chief designers with highly specialised missions. When everything interacts, you need a different structure, and it was part of the genius of Sergei Korolev to create it. (Korolev, who was the author's boss for most of the years described here, is rightly celebrated as a great engineer and champion of missile and space projects, but in Chertok's view at least equally important was his talent in quickly evaluating the potential of individuals and filling jobs with the people [often improbable candidates] best able to do them.)

In this book we see the transformation of the Soviet missile program from slavishly copying German technology to world-class innovation, producing, in short order, the first ICBM, earth satellite, lunar impact, images of the lunar far side, and interplanetary probes. The missile men found themselves vaulted from an obscure adjunct of Red Army artillery to the vanguard of Soviet prestige in the world, with the Soviet leadership urging them on to ever greater exploits.

There is a tremendous amount of detail here—so much that some readers have deemed it tedious: I found it enlightening. The author dissects the Nedelin disaster in forensic detail, as well as the much less known 1980 catastrophe at Plesetsk where 48 died because a component of the rocket used the wrong kind of solder. Rocketry is an exacting business, and it is a gift to generations about to embark upon it to imbibe the wisdom of one who was present at its creation and learned, by decades of experience, just how careful one must be to succeed at it. I could go on regaling you with anecdotes from this book but, hey, if you've made it this far, you're probably going to read it yourself, so what's the point? (But if you do, I'd suggest you read Volume 1 [May 2012] first.)

As with all NASA publications, the work is in the public domain, and an online PDF edition is available.

A Kindle edition is available which is perfectly readable but rather cheaply produced. Footnotes simply appear in the text in-line somewhere after the reference, set in small red type. The index references page numbers from the print edition which are not included in the Kindle version, and hence are completely useless. If you have a workable PDF application on your reading device, I'd go with the NASA PDF, which is not only better formatted but free.

The original Russian edition is available online.

August 2012 Permalink

Chertok, Boris E. Rockets and People. Vol. 3. Washington: National Aeronautics and Space Administration, [1999] 2009. ISBN 978-1-4700-1437-7 NASA SP-2009-4110.
This is the third book of the author's four-volume autobiographical history of the Soviet missile and space program. Boris Chertok was a survivor, living through the Bolshevik revolution, the Russian civil war, Stalin's purges of the 1930s, World War II, all of the postwar conflict between chief designers and their bureaux and rival politicians, and the collapse of the Soviet Union. Born in Poland in 1912, he died in 2011 in Moscow. After retiring from the RKK Energia organisation in 1992 at the age of 80, he wrote this work between 1994 and 1999. Originally published in Russian in 1999, this annotated English translation was prepared by the NASA History Office under the direction of Asif A. Siddiqi, author of Challenge to Apollo (April 2008), the definitive Western history of the Soviet space program.

Volume 2 of this memoir chronicled the achievements which thrust the Soviet Union's missile and space program into the consciousness of people world-wide and sparked the space race with the United States: the development of the R-7 ICBM, Sputnik and its successors, and the first flights which photographed the far side of the Moon and impacted on its surface. In this volume, the author describes the projects and accomplishments which built upon this base and persuaded many observers of the supremacy of Soviet space technology. Since the author's speciality was control systems and radio technology, he had an almost unique perspective upon these events: unlike other designers who focussed upon one or a few projects, he was involved in almost all of the principal efforts, from intermediate range, intercontinental, and submarine-launched ballistic missiles; air and anti-missile defence; piloted spaceflight; reconnaissance, weather, and navigation satellites; communication satellites; deep space missions and the ground support for them; soft landing on the Moon; and automatic rendezvous and docking. He was present when it looked like the rudimentary R-7 ICBM might be launched in anger during the Cuban missile crisis, at the table as chief designers battled over whether combat missiles should use cryogenic or storable liquid propellants or solid fuel, and sat on endless boards of inquiry after mission failures—the first eleven attempts to soft-land on the Moon failed, and Chertok was there for each launch, subsequent tracking, and sorting through what went wrong.

This was a time of triumph for the Soviet space program: the first manned flight, endurance record after endurance record, dual flights, the first woman in space, the first flight with a crew of more than one, and the first spacewalk. But from Chertok's perspective inside the programs, and the freedom he had to write candidly in the 1990s about his experiences, it is clear that the seeds of tragedy were being sown. With the quest for one spectacular after another, each surpassing the last, the Soviets became inoculated with what NASA came to call “go fever”—a willingness to brush anomalies under the rug and normalise the abnormal because you'd gotten away with it before.

One of the most stunning examples of this is Gagarin's flight. The Vostok spacecraft consisted of a spherical descent module (basically a cannonball covered with ablative thermal protection material) and an instrument compartment containing the retro-rocket, attitude control system, and antennas. After firing the retro-rocket, the instrument compartment was supposed to separate, allowing the descent module's heat shield to protect it through atmospheric re-entry. (The Vostok performed a purely ballistic re-entry, and had no attitude control thrusters in the descent module; stability was maintained exclusively by an offset centre of gravity.) In the two unmanned test flights which preceded Garagin's mission, the instrument module had failed to cleanly separate from the descent module, but the connection burned through during re-entry and the descent module survived. Gagarin was launched in a spacecraft with the same design, and the same thing happened: there were wild oscillations, but after the link burned through his spacecraft stabilised. Astonishingly, Vostok 2 was launched with Gherman Titov on board with precisely the same flaw, and suffered the same failure during re-entry. Once again, the cosmonaut won this orbital game of Russian roulette. One wonders what lessons were learned from this. In this narrative, Chertok is simply aghast at the decision making here, but one gets the sense that you had to be there, then, to appreciate what was going through people's heads.

The author was extensively involved in the development of the first Soviet communications satellite, Molniya, and provides extensive insights into its design, testing, and early operations. It is often said that the Molniya orbit was chosen because it made the satellite visible from the Soviet far North where geostationary satellites would be too close to the horizon for reliable communication. It is certainly true that today this orbit continues to be used for communications with Russian arctic territories, but its adoption for the first Soviet communications satellite had an entirely different motivation. Due to the high latitude of the Soviet launch site in Kazakhstan, Korolev's R-7 derived booster could place only about 100 kilograms into a geostationary orbit, which was far too little for a communication satellite with the technology of the time, but it could loft 1,600 kilograms into a high-inclination Molniya orbit. The only alternative would have been for Korolev to have approached Chelomey to launch a geostationary satellite on his UR-500 (Proton) booster, which was unthinkable because at the time the two were bitter rivals. So much for the frictionless efficiency of central planning!

In engineering, one learns that every corner cut will eventually come back to cut you. Korolev died at just the time he was most needed by the Soviet space program due to a botched operation for a routine condition performed by a surgeon who had spent most of his time as a Minister of the Soviet Union and not in the operating room. Gagarin died in a jet fighter training accident which has been the subject of such an extensive and multi-layered cover-up and spin that the author simply cites various accounts and leaves it to the reader to judge. Komarov died in Soyuz 1 due to a parachute problem which would have been discovered had an unmanned flight preceded his. He was a victim of “go fever”.

There is so much insight and wisdom here I cannot possibly summarise it all; you'll have to read this book to fully appreciate it, ideally after having first read Volume 1 (May 2012) and Volume 2 (August 2012). Apart from the unique insider's perspective on the Soviet missile and space program, as a person elected a corresponding member of the Soviet Academy of Sciences in 1968 and a full member (academician) of the Russian Academy of Sciences in 2000, he provides a candid view of the politics of selection of members of the Academy and how they influence policy and projects at the national level. Chertok believes that, even as one who survived Stalin's purges, there were merits to the Soviet system which have been lost in the “new Russia”. His observations are worth pondering by those who instinctively believe the market will always converge upon the optimal solution.

As with all NASA publications, the work is in the public domain, and an online edition in PDF, EPUB, and MOBI formats is available.

A commercial Kindle edition is available which is perfectly readable but rather cheaply produced. Footnotes simply appear in the text in-line somewhere after the reference, set in small red type. The index references page numbers from the print edition which are not included in the Kindle version, and hence are completely useless. If you have a suitable application on your reading device for one of the electronic book formats provided by NASA, I'd opt for it. They are not only better formatted but free.

The original Russian edition is available online.

December 2012 Permalink

Chertok, Boris E. Rockets and People. Vol. 4. Washington: National Aeronautics and Space Administration, [1999] 2011. ISBN 978-1-4700-1437-7 NASA SP-2011-4110.
This is the fourth and final book of the author's autobiographical history of the Soviet missile and space program. Boris Chertok was a survivor, living through the Bolshevik revolution, the Russian civil war, Stalin's purges of the 1930s, World War II, all of the postwar conflict between chief designers and their bureaux and rival politicians, and the collapse of the Soviet Union. Born in Poland in 1912, he died in 2011 in Moscow. As he says in this volume, “I was born in the Russian Empire, grew up in Soviet Russia, achieved a great deal in the Soviet Union, and continue to work in Russia.” After retiring from the RKK Energia organisation in 1992 at the age of 80, he wrote this work between 1994 and 1999. Originally published in Russian in 1999, this annotated English translation was prepared by the NASA History Office under the direction of Asif A. Siddiqi, author of Challenge to Apollo (April 2008), the definitive Western history of the Soviet space program.

This work covers the Soviet manned lunar program and the development of long-duration space stations and orbital rendezvous, docking, and assembly. As always, Chertok was there, and participated in design and testing, was present for launches and in the control centre during flights, and all too often participated in accident investigations.

In retrospect, the Soviet manned lunar program seems almost bizarre. It did not begin in earnest until two years after NASA's Apollo program was underway, and while the Gemini and Apollo programs were a step-by-step process of developing and proving the technologies and operational experience for lunar missions, the Soviet program was a chaotic bag of elements seemingly driven more by the rivalries of the various chief designers than a coherent plan for getting to the Moon. First of all, there were two manned lunar programs, each using entirely different hardware and mission profiles. The Zond program used a modified Soyuz spacecraft launched on a Proton booster, intended to send two cosmonauts on a circumlunar mission. They would simply loop around the Moon and return to Earth without going into orbit. A total of eight of these missions were launched unmanned, and only one completed a flight which would have been safe for cosmonauts on board. After Apollo 8 accomplished a much more ambitious lunar orbital mission in December 1968, a Zond flight would simply demonstrate how far behind the Soviets were, and the program was cancelled in 1970.

The N1-L3 manned lunar landing program was even more curious. In the Apollo program, the choice of mission mode and determination of mass required for the lunar craft came first, and the specifications of the booster rocket followed from that. Work on Korolev's N1 heavy lifter did not get underway until 1965—four years after the Saturn V, and it was envisioned as a general purpose booster for a variety of military and civil space missions. Korolev wanted to use very high thrust kerosene engines on the first stage and hydrogen engines on the upper stages as did the Saturn V, but he was involved in a feud with Valentin Glushko, who championed the use of hypergolic, high boiling point, toxic propellants and refused to work on the engines Korolev requested. Hydrogen propellant technology in the Soviet Union was in its infancy at the time, and Korolev realised that waiting for it to mature would add years to the schedule.

In need of engines, Korolev approached Nikolai Kuznetsov, a celebrated designer of jet turbine engines, but who had no previous experience at all with rocket engines. Kuznetsov's engines were much smaller than Korolev desired, and to obtain the required thrust, required thirty engines on the first stage alone, each with its own turbomachinery and plumbing. Instead of gimballing the engines to change the thrust vector, pairs of engines on opposite sides of the stage were throttled up and down. The gargantuan scale of the lower stages of the N-1 meant they were too large to transport on the Soviet rail network, so fabrication of the rocket was done in a huge assembly hall adjacent to the launch site. A small city had to be built to accommodate the work force.

All Soviet rockets since the R-2 in 1949 had used “integral tanks”: the walls of the propellant tanks were load-bearing and formed the skin of the rocket. The scale of the N1 was such that load-bearing tanks would have required a wall thickness which exceeded the capability of Soviet welding technology at the time, forcing a design with an external load-bearing shell and separate propellant tanks within it. This increased the complexity of the rocket and added dead weight to the design. (NASA's contractors had great difficulty welding the integral tanks of the Saturn V, but NASA simply kept throwing money at the problem until they figured out how to do it.)

The result was a rocket which was simultaneously huge, crude, and bewilderingly complicated. There was neither money in the budget nor time in the schedule to build a test stand to permit ground firings of the first stage. The first time those thirty engines fired up would be on the launch pad. Further, Kuznetsov's engines were not reusable. After every firing, they had to be torn down and overhauled, and hence were essentially a new and untested engine every time they fired. The Saturn V engines, by contrast, while expended in each flight, could be and were individually test fired, then ground tested together installed on the flight stage before being stacked into a launch vehicle.

The weight and less efficient fuel of the N-1 made its performance anæmic. While it had almost 50% more thrust at liftoff than the Saturn V, its payload to low Earth orbit was 25% less. This meant that performing a manned lunar landing mission in a single launch was just barely possible. The architecture would have launched two cosmonauts in a lunar orbital ship. After entering orbit around the Moon, one would spacewalk to the separate lunar landing craft (an internal docking tunnel as used in Apollo would have been too heavy) and descend to the Moon. Fuel constraints meant the cosmonaut only had ten to fifteen seconds to choose a landing spot. After the footprints, flag, and grabbing a few rocks, it was back to the lander to take off to rejoin the orbiter. Then it took another spacewalk to get back inside. Everybody involved at the time was acutely aware how marginal and risky this was, but given that the N-1 design was already frozen and changing it or re-architecting the mission to two or three launches would push out the landing date four or five years, it was the only option that would not forfeit the Moon race to the Americans.

They didn't even get close. In each of its test flights, the N-1 did not even get to the point of second stage ignition (although in its last flight it got within seven seconds of that milestone). On the second test flight the engines cut off shortly after liftoff and the vehicle fell back onto the launch pad, completely obliterating it in the largest artificial non-nuclear explosion known to this date: the equivalent of 7 kilotons of TNT. After four consecutive launch failures, having lost the Moon race, with no other mission requiring its capabilities, and the military opposing an expensive program for which they had no use, work on the N-1 was suspended in 1974 and the program officially cancelled in 1976.

When I read Challenge to Apollo, what struck me was the irony that the Apollo program was the very model of a centrally-planned state-directed effort along Soviet lines, while the Soviet Moon program was full of the kind of squabbling, turf wars, and duplicative competitive efforts which Marxists decry as flaws of the free market. What astounded me in reading this book is that the Soviets were acutely aware of this in 1968. In chapter 9, Chertok recounts a Central Committee meeting in which Minister of Defence Dmitriy Ustinov remarked:

…the Americans have borrowed our basic method of operation—plan-based management and networked schedules. They have passed us in management and planning methods—they announce a launch preparation schedule in advance and strictly adhere to it. In essence, they have put into effect the principle of democratic centralism—free discussion followed by the strictest discipline during implementation.

In addition to the Moon program, there is extensive coverage of the development of automated rendezvous and docking and the long duration orbital station programs (Almaz, Salyut, and Mir). There is also an enlightening discussion, building on Chertok's career focus on control systems, of the challenges in integrating humans and automated systems into the decision loop and coping with off-nominal situations in real time.

I could go on and on, but there is so much to learn from this narrative, I'll just urge you to read it. Even if you are not particularly interested in space, there is much experience and wisdom to be gained from it which are applicable to all kinds of large complex systems, as well as insight into how things were done in the Soviet Union. It's best to read Volume 1 (May 2012), Volume 2 (August 2012), and Volume 3 (December 2012) first, as they will introduce you to the cast of characters and the events which set the stage for those chronicled here.

As with all NASA publications, the work is in the public domain, and an online edition in PDF, EPUB, and MOBI formats is available.

A commercial Kindle edition is available which is much better produced than the Kindle editions of the first three volumes. If you have a suitable application on your reading device for one of the electronic book formats provided by NASA, I'd opt for it. They're free.

The original Russian edition is available online.

March 2013 Permalink

Chiles, Patrick. Perigee. Seattle: CreateSpace, 2011. ISBN 978-1-4699-5713-5.
A few years into the future, while NASA bumbles along in its bureaucratic haze and still can't launch humans into space, a commercial “new space” company, Polaris AeroSpace Lines, has taken the next step beyond suborbital tourist hops into space for the well-heeled, and begun both scheduled and charter service in aerospace planes equipped with a combined-cycle powerplant which allows them to fly anywhere on the globe, operating at Mach 10, making multiple skips off the atmosphere, and delivering up to 30 passengers and cargo to any destination in around 90 minutes. Passengers are treated to a level of service and coddling which exceeds first class, breathtaking views from above the atmosphere along the way, and apart from the steep ticket prices, no downside apart from the zero-g toilet.

In this thriller, something goes horribly wrong during a flight from Denver to Singapore chartered by a coarse and demanding Australian media mogul, and the crew and passengers find themselves not on course for their destination but rather trapped in Earth orbit with no propellant and hence no prospect of getting back until long after their life support will be exhausted. Polaris immediately begins to mount a rescue mission based upon an orbital spacecraft they have under development, but as events play out clues begin to emerge that a series of problems are not systems failures but perhaps evidence of something much darker, in which those on the front lines trying to get their people back do not know who they can trust. Eventually, Polaris has no option but to partner with insurgent individuals in the “old space” world to attempt an improvised rescue mission.

This is a very interesting book, in that it does not read like a space thriller so much as one of the classic aviation dramas such as The High and the Mighty. We have the cast of characters: a crusty mechanic, heroic commander, hot-shot first officer, resourceful flight attendant with unexpected talents, demanding passengers, visionary company president, weaselly subordinate, and square-jawed NASA types. It all works very well, and as long as you don't spend too much time thinking about mass fractions, specific impulse, orbital mechanics, and thermal protection systems, is an enjoyable read, and provides a glimpse of a plausible future for commercial space flight (point to point hypersonic service) which is little discussed among the new space community. For those who do care about the details, they follow. Be warned—some of these are major plot spoilers, so if you're planning to read the novel it's best to give them a pass until you've finished the book.

Spoiler warning: Plot and/or ending details follow.  

  • In chapter 26 we are told that the spaceplane's electricity is produced by fuel cells. This doesn't make any sense for a suborbital craft. We're also told that it is equipped with an APU and batteries with eight hours of capacity. For a plane which can fly to its destination in 90 minutes, why would you also include a fuel cell? The APU can supply power for normal operation, and in case it fails, the batteries have plenty of capacity to get you back on the ground. Also, you'd have to carry liquid hydrogen to power the fuel cells. This would require a bulky tank and make ramp operations and logistics a nightmare.
  • Not a quibble, but rather a belly laugh in chapter 28: I had not before heard the aging International Space Station called “Cattlecar Galactica”.
  • In chapter 31, when the rescue mission is about to launch, we're told that if the launch window is missed, on the next attempt the stricken craft will be “several hundred miles farther downrange”. In fact, the problem is that on the next orbit, due to the Earth's rotation, the plane of the craft's orbit will have shifted with respect to that of the launch site, and consequently the rescue mission will have to perform a plane change as part of its trajectory. This is hideously costly in terms of fuel, and it is unlikely in the extreme the rescue mission would be able to accomplish it. All existing rendezvous missions, if they miss their launch window, must wait until the next day when the launch site once again aligns with the orbital plane of the destination.
  • In chapter 47, as passenger Magrath begins to lose it, “Sweat began to bead up on his bald head and float away.” But in weightlessness, surface tension dominates all other forces and the sweat would cling and spread out over the 'strine's pate. There is nothing to make it float away.
  • In chapter 54 and subsequently, Shuttle “rescue balls” are used to transfer passengers from the crippled spaceplane to the space station. These were said to have been kept on the station since early in the program. In fact, while NASA did develop a prototype of the Personal Rescue Enclosure, they were never flown on any Shuttle mission nor launched to the station.
  • The orbital mechanics make absolutely no sense at all. One would expect a suborbital flight between Denver and Singapore to closely follow a great circle route between those airports (with possible deviations due to noise abatement and other considerations). Since most of the flight would be outside the atmosphere, weather and winds aloft would not be a major consideration. But if flight 501 had followed such a route and have continued to boost into orbit, it would have found itself in a high-inclination retrograde orbit around the Earth: going the opposite direction to the International Space Station. Getting from such an orbit to match orbits with the ISS would require more change in velocity (delta-v) than an orbital launch from the Earth, and no spacecraft in orbit would have remotely that capability. The European service vehicle already docked at the station would only have enough propellant for a destructive re-entry.

    We're told then that the flight path would be to the east, over Europe. but why would one remotely choose such a path, especially if a goal of the flight was to set records? It would be a longer flight, and much more demanding of propellant to do it in one skip as planned. But, OK, let's assume that for some reason they did decide to go the long way around. Now, for the rescue to be plausible, we have to assume two further ridiculously improbable things: first, that the inclination of the orbit resulting from the engine runaway on the flight to Singapore would match that of the station, and second, that the moment of launch just happened to be precisely when Denver was aligned with the plane of the station's orbit. Since there is no reason that the launch would have been scheduled to meet these exacting criteria, the likelihood that the spaceplane would be in an orbit reachable from the station without a large and impossible to accomplish plane change (here, I am referring to a change in the orbital plane, not catching a connecting flight) is negligible.

Spoilers end here.  

The author's career has been in the airline industry, and this shows in the authenticity of the depiction of airline operations. Notwithstanding the natters above behind the spoiler shield, I thoroughly enjoyed this book and raced through it trying to guess how it would come out.

August 2012 Permalink

Chiles, Patrick. Farside. Seattle: Amazon Digital Services, 2015. ASIN B010WAE080.
Several years after the events chronicled in Perigee (August 2012), Arthur Hammond's Polaris AeroSpace Lines is operating routine point-to-point suborbital passenger and freight service with its Clippers, has expanded into orbital service with Block II Clippers, and is on the threshold of opening up service to the Moon with its “cycler” spacecraft which loop continuously between the Earth and Moon. Clippers rendezvous with the cyclers as they approach the Earth, transferring crew, passengers, cargo, and consumables. Initial flights will be limited to lunar orbit, but landing missions are envisioned for the future.

In the first orbital mission, chartered to perform resource exploration from lunar orbit, cycler Shepard is planning to enter orbit with a burn which will, by the necessities of orbital mechanics, have to occur on the far side of the Moon, out of radio contact with the Earth. At Polaris mission control in Denver, there is the usual tension as the clock ticks down toward the time when Shepard is expected to emerge from behind the Moon, safely in orbit. (If the burn did not occur, the ship would appear before this time, still on a trajectory which would return it to the Earth.) When the acquisition of signal time comes and goes with no reply to calls and no telemetry, tension gives way to anxiety. Did Shepard burn too long and crash on the far side of the Moon? Did its engine explode and destroy the ship? Did some type of total system failure completely disable its communications?

On board Shepard, Captain Simon Poole is struggling to survive after the disastrous events which occurred just moments after the start of the lunar orbit insertion burn. Having taken refuge in the small airlock after the expandable habitation module has deflated, he has only meagre emergency rations to sustain him until a rescue mission might reach him. And no way to signal Earth that he is alive.

What seems a terrible situation rapidly gets worse and more enigmatic when an arrogant agent from Homeland Security barges into Polaris and demands information about the passenger and cargo manifest for the flight, Hammond is visited at home by an unlikely caller, and a jarhead/special operator type named Quinn shows them some darker than black intelligence about their ship and “invites” them to NORAD headquarters to be briefed in on an above top secret project.

So begins a nearish future techno-thriller in which the situations are realistic, the characters interesting, the perils harrowing, and the stakes could not be higher. The technologies are all plausible extrapolations of those available at present, with no magic. Government agencies behave as they do in the real world, which is to say with usually good intentions leavened with mediocrity, incompetence, scheming ambition, envy, and counter-productive secrecy and arrogance. This novel is not going to be nominated for any awards by the social justice warriors who have infiltrated the science fiction writer and fan communities: the author understands precisely who the enemies of civilisation and human destiny are, forthrightly embodies them in his villains, and explains why seemingly incompatible ideologies make common cause against the values which have built the modern world. The story is one of problem solving, adventure, survival, improvisation, and includes one of the most unusual episodes of space combat in all of science fiction. It would make a terrific movie.

For the most part, the author gets the details right. There are a few outright goofs, such as seeing the Earth from the lunar far side (where it is always below the horizon—that's why it's the far side); some errors in orbital mechanics which will grate on players of Kerbal Space Program; the deployed B-1B bomber is Mach 1.25, not Mach 2; and I don't think there's any way the ships in the story could have had sufficient delta-v to rendezvous with a comet so far out the plane of the ecliptic. But I'm not going to belabour these quibbles in what is a rip-roaring read. There is a glossary of aerospace terms and acronyms at the end. Also included is a teaser chapter for a forthcoming novel which I can't wait to read.

October 2015 Permalink

Clark, John D. Ignition! New Brunswick, NJ: Rutgers University Press, 1972. ISBN 978-0-8135-0725-5.
This may be the funniest book about chemistry ever written. In the golden age of science fiction, one recurring theme was the search for a super “rocket fuel” (with “fuel” used to mean “propellant”) which would enable the exploits depicted in the stories. In the years between the end of World War II and the winding down of the great space enterprise with the conclusion of the Apollo project, a small band of researchers (no more than 200 in the U.S., of whom around fifty were lead scientists), many of whom had grown up reading golden age science fiction, found themselves tasked to make their boyhood dreams real—to discover exotic propellants which would allow rockets to accomplish missions envisioned not just by visionaries but also the hard headed military men who, for the most part, paid the bills.

Propulsion chemists are a rare and special breed. As Isaac Asimov (who worked with the author during World War II) writes in a short memoir at the start of the book:

Now, it is clear that anyone working with rocket fuels is outstandingly mad. I don't mean garden-variety crazy or merely raving lunatic. I mean a record-shattering exponent of far-out insanity.

There are, after all, some chemicals that explode shatteringly, some that flame ravenously, some that corrode hellishly, some that poison sneakily, and some that stink stenchily. As far as I know, though, only liquid rocket fuels have all these delightful properties combined into one delectable whole.

And yet amazingly, as head of propulsion research at the Naval Air Rocket Test Station and its successor organisation for seventeen years, the author not only managed to emerge with all of his limbs and digits intact, his laboratory never suffered a single time-lost mishap. This, despite routinely working with substances such as:

Chlorine trifluoride, ClF3, or “CTF” as the engineers insist on calling it, is a colorless gas, a greenish liquid, or a white solid. … It is also quite probably the most vigorous fluorinating agent in existence—much more vigorous than fluorine itself. … It is, of course, extremely toxic, but that's the least of the problem. It is hypergolic with every known fuel, and so rapidly hypergolic that no ignition delay has ever been measured. It is also hypergolic with such things as cloth, wood, and test engineers, not to mention asbestos, sand, and water—with which it reacts explosively. It can be kept in some of the ordinary structural metals—steel, copper, aluminum, etc.—because the formation of a thin film of insoluble metal fluoride which protects the bulk of the metal, just as the invisible coat of oxide on aluminum keeps it from burning up in the atmosphere. If, however, this coat is melted or scrubbed off, the operator is confronted with the problem of coping with a metal-fluorine fire. For dealing with this situation, I have always recommended a good pair of running shoes. (p. 73)

And ClF3 is pretty benign compared to some of the other dark corners of chemistry into which their research led them. There is extensive coverage of the quest for a high energy monopropellant, the discovery of which would greatly simplify the design of turbomachinery, injectors, and eliminate problems with differential thermal behaviour and mixture ratio over the operating range of an engine which used it. However, the author reminds us:

A monopropellant is a liquid which contains in itself both the fuel and the oxidizer…. But! Any intimate mixture of a fuel and an oxidizer is a potential explosive, and a molecule with one reducing (fuel) end and one oxidizing end, separated by a pair of firmly crossed fingers, is an invitation to disaster. (p. 10)

One gets an excellent sense of just how empirical all of this was. For example, in the quest for “exotic fuel” (which the author defines as “It's expensive, it's got boron in it, and it probably doesn't work.”), straightforward inorganic chemistry suggested that burning a borane with hydrazine, for example:

2B5H9 + 5N2H4 ⟶ 10BN + 19H2

would be a storable propellant with a specific impulse (Isp) of 326 seconds with a combustion chamber temperature of just 2000°K. But this reaction and the calculation of its performance assumes equilibrium conditions and, apart from a detonation (something else with which propulsion chemists are well acquainted), there are few environments as far from equilibrium as a rocket combustion chamber. In fact, when you try to fire these propellants in an engine, you discover the reaction products actually include elemental boron and ammonia, which result in disappointing performance. Check another one off the list.

Other promising propellants ran afoul of economic considerations and engineering constraints. The lithium, fluorine, and hydrogen tripropellant system has been measured (not theoretically calculated) to have a vacuum Isp of an astonishing 542 seconds at a chamber pressure of only 500 psi and temperature of 2200°K. (By comparison, the space shuttle main engine has a vacuum Isp of 452.3 sec. with a chamber pressure of 2994 psi and temperature of 3588°K; a nuclear thermal rocket would have an Isp in the 850–1000 sec. range. Recall that the relationship between Isp and mass ratio is exponential.) This level of engine performance makes a single stage to orbit vehicle not only feasible but relatively straightforward to engineer. Unfortunately, there is a catch or, to be precise, a list of catches. Lithium and fluorine are both relatively scarce and very expensive in the quantities which would be required to launch from the Earth's surface. They are also famously corrosive and toxic, and then you have to cope with designing an engine in which two of the propellants are cryogenic fluids and the third is a metal which is solid below 180°C. In the end, the performance (which is breathtaking for a chemical rocket) just isn't worth the aggravation.

In the final chapter, the author looks toward the future of liquid rocket propulsion and predicts, entirely correctly from a perspective four decades removed, that chemical propulsion was likely to continue to use the technologies upon which almost all rockets had settled by 1970: LOX/hydrocarbon for large first stages, LOX/LH2 for upper stages, and N2O4/hydrazine for storable missiles and in-space propulsion. In the end economics won out over the potential performance gains to be had from the exotic (and often far too exciting) propellants the author and his colleagues devoted their careers to exploring. He concludes as follows.

There appears to be little left to do in liquid propellant chemistry, and very few important developments to be anticipated. In short, we propellant chemists have worked ourselves out of a job. The heroic age is over.

But it was great fun while it lasted. (p. 192)

Now if you've decided that you just have to read this book and innocently click on the title above to buy a copy, you may be at as much risk of a heart attack as those toiling in the author's laboratory. This book has been out of print for decades and is considered such a classic, both for its unique coverage of the golden age of liquid propellant research, comprehensive description of the many avenues explored and eventually abandoned, hands-on chemist-to-chemist presentation of the motivation for projects and the adventures in synthesising and working with these frisky molecules, not to mention the often laugh out loud writing, that used copies, when they are available, sell for hundreds of dollars. As I am writing these remarks, seven copies are offered at Amazon at prices ranging from US$300–595. Now, this is a superb book, but it isn't that good!

If, however, you type the author's name and the title of the book into an Internet search engine, you will probably quickly come across a PDF edition consisting of scanned pages of the original book. I'm not going to link to it here, both because I don't link to works which violate copyright as a matter of principle and since my linking to a copy of the PDF edition might increase its visibility and risk of being taken down. I am not one of those people who believes “information wants to be free”, but I also doubt John Clark would have wanted his unique memoir and invaluable reference to be priced entirely beyond the means of the vast majority of those who would enjoy and be enlightened by reading it. In the case of “orphaned works”, I believe the moral situation is ambiguous (consider: if you do spend a fortune for a used copy of an out of print book, none of the proceeds benefit the author or publisher in any way). You make the call.

April 2012 Permalink

Dewar, James A. To the End of the Solar System. 2nd. ed. Burlington, Canada: Apogee Books, [2004] 2007. ISBN 978-1-894959-68-1.
If you're seeking evidence that entrusting technology development programs such as space travel to politicians and taxpayer-funded bureaucrats is a really bad idea, this is the book to read. Shortly after controlled nuclear fission was achieved, scientists involved with the Manhattan Project and the postwar atomic energy program realised that a rocket engine using nuclear fission instead of chemical combustion to heat a working fluid of hydrogen would have performance far beyond anything achievable with chemical rockets and could be the key to opening up the solar system to exploration and eventual human settlement. (The key figure of merit for rocket propulsion is “specific impulse”, expressed in seconds, which [for rockets] is simply an odd way of expressing the exhaust velocity. The best chemical rockets have specific impulses of around 450 seconds, while early estimates for solid core nuclear thermal rockets were between 800 and 900 seconds. Note that this does not mean that nuclear rockets were “twice as good” as chemical: because the rocket equation gives the mass ratio [mass of fuelled rocket versus empty mass] as exponential in the specific impulse, doubling that quantity makes an enormous difference in the missions which can be accomplished and drastically reduces the mass which must be lifted from the Earth to mount them.)

Starting in 1955, a project began, initially within the U.S. Air Force and the two main weapons laboratories, Los Alamos and Livermore, to explore near-term nuclear rocket propulsion, initially with the goal of an ICBM able to deliver the massive thermonuclear bombs of the epoch. The science was entirely straightforward: build a nuclear reactor able to operate at a high core temperature, pump liquid hydrogen through it at a large rate, expel the hot gaseous hydrogen through a nozzle, and there's your nuclear rocket. Figure out the temperature of exhaust and the weight of the entire nuclear engine, and you can work out the precise performance and mission capability of the system. The engineering was a another matter entirely. Consider: a modern civil nuclear reactor generates about a gigawatt, and is a massive structure enclosed in a huge containment building with thick radiation shielding. It operates at a temperature of around 300° C, heating pressurised water. The nuclear rocket engine, by comparison, might generate up to five gigawatts of thermal power, with a core operating around 2000° C (compared to the 1132° C melting point of its uranium fuel), in a volume comparable to a 55 gallon drum. In operation, massive quantities of liquid hydrogen (a substance whose bulk properties were little known at the time) would be pumped through the core by a turbopump, which would have to operate in an almost indescribable radiation environment which might flash the hydrogen into foam and would certainly reduce all known lubricants to sludge within seconds. And this was supposed to function for minutes, if not hours (later designs envisioned a 10 hour operating lifetime for the reactor, with 60 restarts after being refuelled for each mission).

But what if it worked? Well, that would throw open the door to the solar system. Instead of absurd, multi-hundred-billion dollar Mars programs that land a few civil servant spacemen for footprints, photos, and a few rocks returned, you'd end up, for an ongoing budget comparable to that of today's grotesque NASA jobs program, with colonies on the Moon and Mars working their way toward self-sufficiency, regular exploration of the outer planets and moons with mission durations of years, not decades, and the ability to permanently expand the human presence off this planet and simultaneously defend the planet and its biosphere against the kind of Really Bad Day that did in the dinosaurs (and a heck of a lot of other species nobody ever seems to mention).

Between 1955 and 1973, the United States funded a series of projects, usually designated as Rover and NERVA, with the potential of achieving all of this. This book is a thoroughly documented (65 pages of end notes) and comprehensive narrative of what went wrong. As is usually the case when government gets involved, almost none of the problems were technological. The battles, and the eventual defeat of the nuclear rocket were due to agencies fighting for turf, bureaucrats seeking to build their careers by backing or killing a project, politicians vying to bring home the bacon for their constituents or kill projects of their political opponents, and the struggle between the executive and legislative branches and the military for control over spending priorities.

What never happened among all of the struggles and ups and downs documented here is an actual public debate over the central rationale of the nuclear rocket: should there be, or should there not be, an expansive program (funded within available discretionary resources) to explore, exploit the resources, and settle the solar system? Because if no such program were contemplated, then a nuclear rocket would not be required and funds spent on it squandered. But if such a program were envisioned and deemed worthy of funding, a nuclear rocket, if feasible, would reduce the cost and increase the capability of the program to such an extent that the research and development cost of nuclear propulsion would be recouped shortly after the resulting technology were deployed.

But that debate was never held. Instead, the nuclear rocket program was a political football which bounced around for 18 years, consuming 1.4 billion (p. 207) then-year dollars (something like 5.3 billion in today's incredible shrinking greenbacks). Goals were redefined, milestones changed, management shaken up and reorganised, all at the behest of politicians, yet through it all virtually every single technical goal was achieved on time and often well ahead of schedule. Indeed, when the ball finally bounced out of bounds and the 8000 person staff was laid off, dispersing forever their knowledge of the “black art” of fuel element, thermal, and neutronic design constraints for such an extreme reactor, it was not because the project was judged infeasible, but the opposite. The green eyeshade brigade considered the project too likely to succeed, and feared the funding requests for the missions which this breakthrough technological capability would enable. And so ended the possibility of human migration into the solar system for my generation. So it goes. When the rock comes down, the few transient survivors off-planet will perhaps recall their names; they are documented here.

There are many things to criticise about this book. It is cheaply made: the text is set in painfully long lines in a small font with narrow margins, which require milliarcsecond-calibrated eye muscles to track from the end of a line to the start of the next. The printing lops off the descenders from the last line of many pages, leaving the reader to puzzle over words like “hvdrooen” and phrases such as “Whv not now?”. The cover seems to incorporate some proprietary substance made of kangaroo hair and discarded slinkies which makes it curl into a tube once you've opened it and read a few pages. Now, these are quibbles which do not detract from the content, but then this is a 300 page paperback without a single colour plate with a cover price of USD26.95. There are a number of factual errors in the text, but none which seriously distort the meaning for the knowledgeable reader; there are few, if any, typographical errors. The author is clearly an enthusiast for nuclear rocket technology, and this sometimes results in over-the-top hyperbole where a dispassionate recounting of the details should suffice. He is a big fan of New Mexico senator Clinton Anderson, a stalwart supporter of the nuclear rocket from its inception through its demise (which coincided with his retirement from the Senate due to health reasons), but only on p. 145 does the author address the detail that the programme was a multi-billion dollar (in an epoch when a billion dollars was real money) pork barrel project for Anderson's state.

Flawed—yes, but if you're interested in this little-known backstory of the space program of the last century, whose tawdry history and shameful demise largely explains the sorry state of the human presence in space today, this is the best source of which I'm aware to learn what happened and why. Given the cognitive collapse in the United States (Want to clear a room of Americans? Just say “nuclear!”), I can't share the author's technologically deterministic optimism, “The potential foretells a resurgence at Jackass Flats…” (p. 195), that the legacy of Rover/NERVA will be redeemed by the descendants of those who paid for it only to see it discarded. But those who use this largely forgotten and, in the demographically imploding West, forbidden knowledge to make the leap off our planet toward our destiny in the stars will find the experience summarised here, and the sources cited, an essential starting point for the technologies they'll require to get there.

 ‘Und I'm learning Chinese,’ says Wernher von Braun.

June 2008 Permalink

Dewar, James with Robert Bussard. The Nuclear Rocket. Burlington, Canada: Apogee Books, 2009. ISBN 978-1-894959-99-5.
Let me begin with a few comments about the author attribution of this book. I have cited it as given on the copyright page, but as James Dewar notes in his preface, the main text of the book is entirely his creation. He says of Robert Bussard, “I am deeply indebted to Bob's contributions and consequently list his name in the credit to this book”. Bussard himself contributes a five-page introduction in which he uses, inter alia, the adjectives “amazing”, “strange”, “remarkable”, “wonderful”, “visionary”, and “most odd” to describe the work, which he makes clear is entirely Dewar's. Consequently, I shall subsequently use “the author” to denote Dewar alone. Bussard died in 2007, two years before the publication of this book, so his introduction must have been based upon a manuscript. I leave to the reader to judge the propriety of posthumously naming as co-author a prominent individual who did not write a single word of the main text.

Unlike the author's earlier To the End of the Solar System (June 2008), which was a nuts and bolts history of the U.S. nuclear rocket program, this book, titled The Nuclear Rocket, quoting from Bussard's introduction, “…is not really about nuclear rocket propulsion or its applications to space flight…”. Indeed, although some of the nitty-gritty of nuclear rocket engines are discussed, the bulk of the book is an argument for a highly-specific long term plan to transform human access to space from an elitist government run program to a market-driven expansive program with the ultimate goal of providing access to space to all and opening the solar system to human expansion and eventual dominion. This is indeed ambitious and visionary, but of all of Bussard's adjectives, the one that sticks with me is “most odd”.

Dewar argues that the NERVA B-4 nuclear thermal rocket core, developed between 1960 and 1972, and successfully tested on several occasions, has the capability, once the “taboo” against using nuclear engines in the boost to low Earth orbit (LEO) is discarded, of revolutionising space transportation and so drastically reducing the cost per unit mass to orbit that it would effectively democratise access to space. In particular, he proposes a “Re-core” engine which, integrated with a liquid hydrogen tank and solid rocket boosters, would be air-launched from a large cargo aircraft such as a C-5, with the solid rockets boosting the nuclear engine to around 30 km where they would separate for recovery and the nuclear engine engaged. The nuclear rocket would continue to boost the payload to orbital insertion. Since the nuclear stage would not go critical until having reached the upper atmosphere, there would be no radioactivity risk to those handling the stage on the ground prior to launch or to the crew of the plane which deployed the rocket.

After reaching orbit, the payload and hydrogen tank would be separated, and the nuclear engine enclosed in a cocoon (much like an ICBM reentry vehicle) which would de-orbit and eventually land at sea in a region far from inhabited land. The cocoon, which would float after landing, would be recovered by a ship, placed in a radiation-proof cask, and returned to a reprocessing centre where the highly radioactive nuclear fuel core would be removed for reprocessing (the entire launch to orbit would consume only about 1% of the highly enriched uranium in the core, so recovering the remaining uranium and reusing it is essential to the economic viability of the scheme). Meanwhile, another never critical core would be inserted in the engine which, after inspection of the non-nuclear components, would be ready for another flight. If each engine were reused 100 times, and efficient fuel reprocessing were able to produce new cores economically, the cost for each 17,000 pound payload to LEO would be around US$108 per pound.

Payloads which reached LEO and needed to go beyond (for example, to geostationary orbit, the Moon, or the planets) would rendezvous with a different variant of the NERVA-derived engine, dubbed the “Re-use” stage, which is much like Von Braun's nuclear shuttle concept. This engine, like the original NERVA, would be designed for multiple missions, needing only inspection and refuelling with liquid hydrogen. A single Re-use stage might complete 30 round-trip missions before being disposed of in deep space (offering “free launches” for planetary science missions on its final trip into the darkness).

There is little doubt that something like this is technically feasible. After all, the nuclear rocket engine was extensively tested in the years prior to its cancellation in 1972, and NASA's massive resources of the epoch examined mission profiles (under the constraint that nuclear engines could be used only for departure from LEO, however, and without return to Earth) and found no show stoppers. Indeed, there is evidence that the nuclear engine was cancelled, in part, because it was performing so well that policy makers feared it would enable additional costly NASA missions post-Apollo. There are some technological issues: for example, the author implies that the recovered Re-core, once its hot core is extracted and a new pure uranium core installed, will not be radioactive and hence safe to handle without special precautions. But what about neutron activation of other components of the engine? An operating nuclear rocket creates one of the most extreme neutronic environments outside the detonation of a nuclear weapon. Would it be possible to choose materials for the non-core components of the engine which would be immune to this and, if not, how serious would the induced radioactivity be, especially if the engine were reused up to a hundred times? The book is silent on this and a number of other questions.

The initial breakthrough in space propulsion from the first generation nuclear engines is projected to lead to rapid progress in optimising them, with four generations of successively improved engines within a decade or so. This would eventually lead to the development of a heavy lifter able to orbit around 150,000 pounds of payload per flight at a cost (after development costs are amortised or expensed) of about US$87 per pound. This lifter would allow the construction of large space stations and the transport of people to them in “buses” with up to thirty passengers per mission. Beyond that, a nuclear single stage to orbit vehicle is examined, but there are a multitude of technological and policy questions to be resolved before that could be contemplated.

All of this, however, is not what the book is about. The author is a passionate believer in the proposition that opening the space frontier to all the people of Earth, not just a few elite civil servants, is essential to preserving peace, restoring the optimism of our species, and protecting the thin biosphere of this big rock we inhabit. And so he proposes a detailed structure for accomplishing these goals, beginning with “Democratization of Space Act” to be adopted by the U.S. Congress, and the creation of a “Nuclear Rocket Development and Operations Corporation” (NucRocCorp), which would be a kind of private/public partnership in which individuals could invest. This company could create divisions (in some cases competing with one another) and charter development projects. It would entirely control space nuclear propulsion, with oversight by U.S. government regulatory agencies, which would retain strict control over the fissile reactor cores.

As the initial program migrated to the heavy lifter, this structure would morph into a multinational (admitting only “good” nations, however) structure of bewildering (to this engineer) bureaucratic complexity which makes the United Nations look like the student council of Weemawee High. The lines of responsibility and power here are diffuse in the extreme. Let me simply cite “The Stockholder's Declaration” from p. 161:

Whoever invests in the NucRocCorp and subsequent Space Charter Authority should be required to sign a declaration that commits him or her to respect the purpose of the new regime, and conduct their personal lives in a manner that recognizes the rights of their fellow man (What about woman?—JW). They must be made aware that failure to do so could result in forfeiture of their investment.

Property rights, anybody? Thought police? Apart from the manifest baroque complexity of the proposed scheme, it entirely ignores Jerry Pournelle's Iron Law of Bureaucracy: regardless of its original mission, any bureaucracy will eventually be predominately populated by those seeking to advance the interests of the bureaucracy itself, not the purpose for which it was created. The structure proposed here, even if enacted (implausible in the extreme) and even if it worked as intended (vanishingly improbable), would inevitably be captured by the Iron Law and become something like, well, NASA.

On pp. 36–37, the author likens attempts to stretch chemical rocket technology to its limits to gold plating a nail when what is needed is a bigger hammer (nuclear rockets). But this book brings to my mind another epigram: “When all you have is a hammer, everything looks like a nail.” Dewar passionately supports nuclear rocket technology and believes that it is the way to open the solar system to human settlement. I entirely concur. But when it comes to assuming that boosting people up to a space station (p. 111):

And looking down on the bright Earth and into the black heavens might create a new perspective among Protestant, Roman Catholic, and Orthodox theologians, and perhaps lead to the end of the schism plaguing Christianity. The same might be said of the division between the Sunnis and Shiites in Islam, and the religions of the Near and Far East might benefit from a new perspective.

Call me cynical, but I'll wager this particular swing of the hammer is more likely to land on a thumb than the intended nail. Those who cherish individual freedom have often dreamt of a future in which the opening of access to space would, in the words of L. Neil Smith, extend the human prospect to “freedom, immortality, and the stars”—works for me. What is proposed here, if adopted, looks more like, after more than a third of a century of dithering, the space frontier being finally opened to the brave pioneers ready to homestead there, and when they arrive, the tax man and the all-pervasive regulatory state are already there, up and running. The nuclear rocket can expand the human presence throughout the solar system. Let's just hope that when humanity (or some risk-taking subset of it) takes that long-deferred step, it does not propagate the soft tyranny of present day terrestrial governance to worlds beyond.

October 2009 Permalink

Doran, Jamie and Piers Bizony. Starman: the Truth Behind the Legend of Yuri Gagarin. London: Bloomsbury, 1998. ISBN 0-7475-3688-0.

January 2001 Permalink

Dornberger, Walter. V-2. Translated by James Cleugh and Geoffrey Halliday. New York: Ballantine Books, [1952] 1954. LCCN 54-007830.
This book has been out of print for more than forty years. Used copies are generally available via abebooks.com, but the original Viking Press hardcover can be quite expensive. It's wisest to opt for the mass-market Ballantine paperback reprint; copies in perfectly readable condition can usually be had for about US$5.

Dornberger's account is an insider's view of Peenemünde. For an historical treatment with more technical detail plus a description of postwar research using the V-2, see Ordway and Sharpe's 1979 The Rocket Team, ISBN 0-262-65013-4, also out of print but readily available used.

December 2002 Permalink

Dyson, George. Project Orion: The True Story of the Atomic Spaceship. New York: Henry Holt, 2002. ISBN 0-8050-5985-7.

July 2002 Permalink

Easton, Richard D. and Eric F. Frazier. GPS Declassified. Lincoln, NE: Potomac Books, 2013. ISBN 978-1-61234-408-9.
At the dawn of the space age, as the United States planned to launch its Vanguard satellites during the International Geophysical Year (1957–1958), the need to track the orbit of the satellites became apparent. Optical and radar tracking were considered (and eventually used for various applications), but for the first very small satellites would have been difficult. The Naval Research Laboratory proposed a system, Minitrack, which would use the radio beacon of the satellite, received by multiple ground stations on the Earth, which by interferometry would determine the position and velocity of a satellite with great precision. For the scheme to work, a “fence” of receiving stations would have to be laid out which the satellite would regularly cross in its orbit, the positions of each of the receiving stations would have to be known very accurately, and clocks at all of the receiving stations would have to be precisely synchronised with a master clock at the control station which calculated the satellite's orbit.

The technical challenges were overcome, and Minitrack stations were placed into operation at locations within the United States and as far flung as Cuba, Panama, Ecuador, Peru, Chile, Australia, and in the Caribbean. Although designed to track the U.S. Vanguard satellites, after the unexpected launch of Sputnik, receivers were hastily modified to receive the frequency on which it transmitted its beeps, and the system successfully proved itself tracking the first Earth satellite. Minitrack was used to track subsequent U.S. and Soviet satellites until it was supplanted in 1962 by the more capable Spacecraft Tracking and Data Acquisition Network.

An important part of creative engineering is discovering that once you've solved one problem, you may now have the tools at hand to address other tasks, sometimes more important that the one which motivated the development of the enabling technologies in the first place. It didn't take long for a group of engineers at the Naval Research Laboratory (NRL) to realise that if you could determine the precise position and velocity of a satellite in orbit by receiving signals simultaneously at multiple stations on the ground with precisely-synchronised clocks, you could invert the problem and, by receiving signals from multiple satellites in known orbits, each with an accurate and synchronised clock on board, it would be possible to determine the position, altitude, and velocity of the receiver on or above the Earth (and, in addition, provide a precise time signal). With a sufficiently extensive constellation of satellites, precision navigation and time signals could be extended to the entire planet. This was the genesis of the Global Positioning System (GPS) which has become a ubiquitous part of our lives today.

At the start, this concept was “exploratory engineering”: envisioning what could be done (violating no known law of physics) if and when technology advanced to a stage which permitted it. The timing accuracy required for precision navigation could be achieved by atomic clocks (quartz frequency standards were insufficiently stable and subject to drift due to temperature, pressure, and age of the crystal), but in the 1950s and early '60s, atomic clocks were large, heavy, and delicate laboratory apparatus which nobody imagined could be put on top of a rocket and shot into Earth orbit. Just launching single satellites into low Earth orbit was a challenge, with dramatic launch failures and in-orbit malfunctions all too common. The thought of operating a constellation of dozens of satellites in precisely-specified high orbits seemed like science fiction. And even if the satellites with atomic clocks could somehow be launched, the radio technology to receive the faint signals from space and computation required to extract position and velocity information from the signal was something which might take a room full of equipment: hardly practical for a large aircraft or even a small ship.

But the funny thing about an exponentially growing technology is if something seems completely infeasible today, just wait a few years. Often, it will move from impossible to difficult to practical for limited applications to something in everybody's pocket. So it has been with GPS, as this excellent book recounts. In 1964, engineers at NRL (including author Easton's father, Roger L. Easton) proposed a system called Timation, in which miniaturised and ruggedised atomic clocks on board satellites would provide time signals which could be used for navigation on land, sea, and air. After ground based tests and using aircraft to simulate the satellite signal, in 1967 the Timation I satellite was launched to demonstrate the operation of an atomic clock in orbit and use of its signals on the ground. With a single satellite in a relatively low orbit, the satellite would only be visible from a given location for thirteen minutes at a time, but this was sufficient to demonstrate the feasibility of the concept.

As the Timation concept was evolving (a second satellite test was launched in 1969, demonstrating improved accuracy), it was not without competition. The U.S. had long been operating the LORAN system for coarse-grained marine and aircraft navigation, and had beacons marking airways across the country. Starting in 1964, the U.S. Navy's Transit satellite navigation system (which used a Doppler measurement system and did not require a precise clock on the satellites) provided periodic position fixes for Navy submarines and surface ships, but was inadequate for aircraft navigation. In the search for a more capable system, Timation competed with an Air Force proposal for regional satellite constellations including geosynchronous and inclined elliptical orbit satellites.

The development of GPS began in earnest in 1973, with the Air Force designated as the lead service. This project launch occurred in the midst of an inter-service rivalry over navigation systems which did not abate with the official launch of the project. Indeed, even in retrospect, participants in the program dispute how much the eventually deployed system owes to its various precursors. Throughout the 1970s the design of the system was refined and pathfinder technology development missions launched, with the first launch of an experimental satellite in February 1978. One satellite is a stunt, but by 1985 a constellation of 10 experimental satellites were in orbit, allowing the performance of the system to be evaluated, constellation management tools to be developed and tested, and receiver hardware to be checked out. Starting in 1989 operational satellites began to be launched, but it was not until 1993 that worldwide, round-the clock coverage was available, and the high-precision military signal was not declared operational until 1995.

Even though GPS coverage was spotty and not continuous, GPS played an important part in the first Gulf War of 1990–1991. Because the military had lagged in procuring GPS receivers for the troops, large numbers of commercial GPS units were purchased and pressed into service for navigating in the desert. A few GPS-guided weapons were used in the conflict, but their importance was insignificant compared to other precision-guided munitions.

Prior to May 2000 the civilian GPS signal was deliberately degraded in accuracy (can't allow the taxpayers who paid for it to have the same quality of navigation as costumed minions of the state!) This so-called “selective availability” was finally discontinued, making GPS practical for vehicle and non-precision air navigation. GPS units began to appear on the consumer market, and like other electronic gadgets got smaller, lighter, less expensive, and more capable with every passing year. Adoption of GPS for tracking of fleets of trucks, marine navigation, and aircraft use became widespread.

Now that GPS is commonplace and hundreds of millions of people are walking around with GPS receivers in their smartphones, there is a great deal of misunderstanding about precisely what GPS entails. GPS—the Global Positioning System—is precisely that: a system which allows anybody with a compatible receiver and a view of the sky which allows them to see four or more satellites to determine their state vector (latitude, longitude, and altitude, plus velocity in each of those three directions) in a specified co-ordinate system (where much additional complexity lurks, which I'll gloss over here), along with the precise time of the measurement. That's all it does. GPS is entirely passive: the GPS receiver sends nothing back to the satellite, and hence the satellite system is able to accommodate an unlimited number of GPS receivers simultaneously. There is no such thing as a “GPS tracker” which can monitor the position of something via satellite. Trackers use GPS to determine their position, but then report the position by other means (for example, the mobile phone network). When people speak of “their GPS” giving directions, GPS is only telling them where they are and where they're going at each instant. All the rest: map display, turn-by-turn directions, etc. is a “big data” application running either locally on the GPS receiver or using resources in the “cloud”: GPS itself plays no part in this (and shouldn't be blamed when “your GPS” sends you the wrong way down a one-way street).

So successful has GPS been, and so deeply has it become embedded in our technological society and economy, that there are legitimate worries about such a system being under the sole control of the U.S. Air Force which could, if ordered, shut down the civilian GPS signals worldwide or regionally (because of the altitude of the satellites, fine-grained denial of GPS availability would not be possible). Also, the U.S. does not have the best record of maintaining vital infrastructure and has often depended upon weather satellites well beyond their expected lifetimes due to budget crunches. Consequently, other players have entered the global positioning market, with the Soviet/Russian GLONASS, European Galileo, and Chinese BeiDou systems operational or under construction. Other countries, including Japan, India, and Iran, are said to be developing their own regional navigation systems. So far, cooperation among these operators has been relatively smooth, reducing the likelihood of interference and making it possible for future receivers to use multiple constellations for better coverage and precision.

This is a comprehensive history of navigation systems and GPS from inception to the present day, with a look into the future. Extensive source citations are given (almost 40% of the book is end notes), and in the Kindle edition the notes, Web documents cited within them, and the index are all properly linked. There are abundant technical details about the design and operation of the system, but the book is entirely accessible to the intelligent layman. In the lifetimes of all but the youngest people on Earth, GPS has transformed our world into a place where nobody need ever be lost. We are just beginning to see the ramifications of this technology on the economy and how we live our day-to-day lives (for example, the emerging technology of self-driving cars would be impossible without GPS). This book is an essential history of how this technology came to be, how it works, and where it may be going in the future.

July 2015 Permalink

Eyles, Don. Sunburst and Luminary. Boston: Fort Point Press, 2018. ISBN 978-0-9863859-3-3.
In 1966, the author graduated from Boston University with a bachelor's degree in mathematics. He had no immediate job prospects or career plans. He thought he might be interested in computer programming due to a love of solving puzzles, but he had never programmed a computer. When asked, in one of numerous job interviews, how he would go about writing a program to alphabetise a list of names, he admitted he had no idea. One day, walking home from yet another interview, he passed an unimpressive brick building with a sign identifying it as the “MIT Instrumentation Laboratory”. He'd heard a little about the place and, on a lark, walked in and asked if they were hiring. The receptionist handed him a long application form, which he filled out, and was then immediately sent to interview with a personnel officer. Eyles was amazed when the personnel man seemed bent on persuading him to come to work at the Lab. After reference checking, he was offered a choice of two jobs: one in the “analysis group” (whatever that was), and another on the team developing computer software for landing the Apollo Lunar Module (LM) on the Moon. That sounded interesting, and the job had another benefit attractive to a 21 year old just graduating from university: it came with deferment from the military draft, which was going into high gear as U.S. involvement in Vietnam deepened.

Near the start of the Apollo project, MIT's Instrumentation Laboratory, led by the legendary “Doc” Charles Stark Draper, won a sole source contract to design and program the guidance system for the Apollo spacecraft, which came to be known as the “Apollo Primary Guidance, Navigation, and Control System” (PGNCS, pronounced “pings”). Draper and his laboratory had pioneered inertial guidance systems for aircraft, guided missiles, and submarines, and had in-depth expertise in all aspects of the challenging problem of enabling the Apollo spacecraft to navigate from the Earth to the Moon, land on the Moon, and return to the Earth without any assistance from ground-based assets. In a normal mission, it was expected that ground-based tracking and computers would assist those on board the spacecraft, but in the interest of reliability and redundancy it was required that completely autonomous navigation would permit accomplishing the mission.

The Instrumentation Laboratory developed an integrated system composed of an inertial measurement unit consisting of gyroscopes and accelerometers that provided a stable reference from which the spacecraft's orientation and velocity could be determined, an optical telescope which allowed aligning the inertial platform by taking sightings on fixed stars, and an Apollo Guidance Computer (AGC), a general purpose digital computer which interfaced to the guidance system, thrusters and engines on the spacecraft, the astronauts' flight controls, and mission control, and was able to perform the complex calculations for en route maneuvers and the unforgiving lunar landing process in real time.

Every Apollo lunar landing mission carried two AGCs: one in the Command Module and another in the Lunar Module. The computer hardware, basic operating system, and navigation support software were identical, but the mission software was customised due to the different hardware and flight profiles of the Command and Lunar Modules. (The commonality of the two computers proved essential in getting the crew of Apollo 13 safely back to Earth after an explosion in the Service Module cut power to the Command Module and disabled its computer. The Lunar Module's AGC was able to perform the critical navigation and guidance operations to put the spacecraft back on course for an Earth landing.)

By the time Don Eyles was hired in 1966, the hardware design of the AGC was largely complete (although a revision, called Block II, was underway which would increase memory capacity and add some instructions which had been found desirable during the initial software development process), the low-level operating system and support libraries (implementing such functionality as fixed point arithmetic, vector, and matrix computations), and a substantial part of the software for the Command Module had been written. But the software for actually landing on the Moon, which would run in the Lunar Module's AGC, was largely just a concept in the minds of its designers. Turning this into hard code would be the job of Don Eyles, who had never written a line of code in his life, and his colleagues. They seemed undaunted by the challenge: after all, nobody knew how to land on the Moon, so whoever attempted the task would have to make it up as they went along, and they had access, in the Instrumentation Laboratory, to the world's most experienced team in the area of inertial guidance.

Today's programmers may be amazed it was possible to get anything at all done on a machine with the capabilities of the Apollo Guidance Computer, no less fly to the Moon and land there. The AGC had a total of 36,864 15-bit words of read-only core rope memory, in which every bit was hand-woven to the specifications of the programmers. As read-only memory, the contents were completely fixed: if a change was required, the memory module in question (which was “potted” in a plastic compound) had to be discarded and a new one woven from scratch. There was no way to make “software patches”. Read-write storage was limited to 2048 15-bit words of magnetic core memory. The read-write memory was non-volatile: its contents were preserved across power loss and restoration. (Each memory word was actually 16 bits in length, but one bit was used for parity checking to detect errors and not accessible to the programmer.) Memory cycle time was 11.72 microseconds. There was no external bulk storage of any kind (disc, tape, etc.): everything had to be done with the read-only and read-write memory built into the computer.

The AGC software was an example of “real-time programming”, a discipline with which few contemporary programmers are acquainted. As opposed to an “app” which interacts with a user and whose only constraint on how long it takes to respond to requests is the user's patience, a real-time program has to meet inflexible constraints in the real world set by the laws of physics, with failure often resulting in disaster just as surely as hardware malfunctions. For example, when the Lunar Module is descending toward the lunar surface, burning its descent engine to brake toward a smooth touchdown, the LM is perched atop the thrust vector of the engine just like a pencil balanced on the tip of your finger: it is inherently unstable, and only constant corrections will keep it from tumbling over and crashing into the surface, which would be bad. To prevent this, the Lunar Module's AGC runs a piece of software called the digital autopilot (DAP) which, every tenth of a second, issues commands to steer the descent engine's nozzle to keep the Lunar Module pointed flamy side down and adjusts the thrust to maintain the desired descent velocity (the thrust must be constantly adjusted because as propellant is burned, the mass of the LM decreases, and less thrust is needed to maintain the same rate of descent). The AGC/DAP absolutely must compute these steering and throttle commands and send them to the engine every tenth of a second. If it doesn't, the Lunar Module will crash. That's what real-time computing is all about: the computer has to deliver those results in real time, as the clock ticks, and if it doesn't (for example, it decides to give up and flash a Blue Screen of Death instead), then the consequences are not an irritated or enraged user, but actual death in the real world. Similarly, every two seconds the computer must read the spacecraft's position from the inertial measurement unit. If it fails to do so, it will hopelessly lose track of which way it's pointed and how fast it is going. Real-time programmers live under these demanding constraints and, especially given the limitations of a computer such as the AGC, must deploy all of their cleverness to meet them without fail, whatever happens, including transient power failures, flaky readings from instruments, user errors, and completely unanticipated “unknown unknowns”.

The software which ran in the Lunar Module AGCs for Apollo lunar landing missions was called LUMINARY, and in its final form (version 210) used on Apollo 15, 16, and 17, consisted of around 36,000 lines of code (a mix of assembly language and interpretive code which implemented high-level operations), of which Don Eyles wrote in excess of 2,200 lines, responsible for the lunar landing from the start of braking from lunar orbit through touchdown on the Moon. This was by far the most dynamic phase of an Apollo mission, and the most demanding on the limited resources of the AGC, which was pushed to around 90% of its capacity during the final landing phase where the astronauts were selecting the landing spot and guiding the Lunar Module toward a touchdown. The margin was razor-thin, and that's assuming everything went as planned. But this was not always the case.

It was when the unexpected happened that the genius of the AGC software and its ability to make the most of the severely limited resources at its disposal became apparent. As Apollo 11 approached the lunar surface, a series of five program alarms: codes 1201 and 1202, interrupted the display of altitude and vertical velocity being monitored by Buzz Aldrin and read off to guide Neil Armstrong in flying to the landing spot. These codes both indicated out-of-memory conditions in the AGC's scarce read-write memory. The 1201 alarm was issued when all five of the 44-word vector accumulator (VAC) areas were in use when another program requested to use one, and 1202 signalled exhaustion of the eight 12-word core sets required by each running job. The computer had a single processor and could execute only one task at a time, but its operating system allowed lower priority tasks to be interrupted in order to service higher priority ones, such as the time-critical autopilot function and reading the inertial platform every two seconds. Each suspended lower-priority job used up a core set and, if it employed the interpretive mathematics library, a VAC, so exhaustion of these resources usually meant the computer was trying to do too many things at once. Task priorities were assigned so the most critical functions would be completed on time, but computer overload signalled something seriously wrong—a condition in which it was impossible to guarantee all essential work was getting done.

In this case, the computer would throw up its hands, issue a program alarm, and restart. But this couldn't be a lengthy reboot like customers of personal computers with millions of times the AGC's capacity tolerate half a century later. The critical tasks in the AGC's software incorporated restart protection, in which they would frequently checkpoint their current state, permitting them to resume almost instantaneously after a restart. Programmers estimated around 4% of the AGC's program memory was devoted to restart protection, and some questioned its worth. On Apollo 11, it would save the landing mission.

Shortly after the Lunar Module's landing radar locked onto the lunar surface, Aldrin keyed in the code to monitor its readings and immediately received a 1202 alarm: no core sets to run a task; the AGC restarted. On the communications link Armstrong called out “It's a 1202.” and Aldrin confirmed “1202.”. This was followed by fifteen seconds of silence on the “air to ground” loop, after which Armstrong broke in with “Give us a reading on the 1202 Program alarm.” At this point, neither the astronauts nor the support team in Houston had any idea what a 1202 alarm was or what it might mean for the mission. But the nefarious simulation supervisors had cranked in such “impossible” alarms in earlier training sessions, and controllers had developed a rule that if an alarm was infrequent and the Lunar Module appeared to be flying normally, it was not a reason to abort the descent.

At the Instrumentation Laboratory in Cambridge, Massachusetts, Don Eyles and his colleagues knew precisely what a 1202 was and found it was deeply disturbing. The AGC software had been carefully designed to maintain a 10% safety margin under the worst case conditions of a lunar landing, and 1202 alarms had never occurred in any of their thousands of simulator runs using the same AGC hardware, software, and sensors as Apollo 11's Lunar Module. Don Eyles' analysis, in real time, just after a second 1202 alarm occurred thirty seconds later, was:

Again our computations have been flushed and the LM is still flying. In Cambridge someone says, “Something is stealing time.” … Some dreadful thing is active in our computer and we do not know what it is or what it will do next. Unlike Garman [AGC support engineer for Mission Control] in Houston I know too much. If it were in my hands, I would call an abort.

As the Lunar Module passed 3000 feet, another alarm, this time a 1201—VAC areas exhausted—flashed. This is another indication of overload, but of a different kind. Mission control immediately calls up “We're go. Same type. We're go.” Well, it wasn't the same type, but they decided to press on. Descending through 2000 feet, the DSKY (computer display and keyboard) goes blank and stays blank for ten agonising seconds. Seventeen seconds later another 1202 alarm, and a blank display for two seconds—Armstrong's heart rate reaches 150. A total of five program alarms and resets had occurred in the final minutes of landing. But why? And could the computer be trusted to fly the return from the Moon's surface to rendezvous with the Command Module?

While the Lunar Module was still on the lunar surface Instrumentation Laboratory engineer George Silver figured out what happened. During the landing, the Lunar Module's rendezvous radar (used only during return to the Command Module) was powered on and set to a position where its reference timing signal came from an internal clock rather than the AGC's master timing reference. If these clocks were in a worst case out of phase condition, the rendezvous radar would flood the AGC with what we used to call “nonsense interrupts” back in the day, at a rate of 800 per second, each consuming one 11.72 microsecond memory cycle. This imposed an additional load of more than 13% on the AGC, which pushed it over the edge and caused tasks deemed non-critical (such as updating the DSKY) not to be completed on time, resulting in the program alarms and restarts. The fix was simple: don't enable the rendezvous radar until you need it, and when you do, put the switch in the position that synchronises it with the AGC's clock. But the AGC had proved its excellence as a real-time system: in the face of unexpected and unknown external perturbations it had completed the mission flawlessly, while alerting its developers to a problem which required their attention.

The creativity of the AGC software developers and the merit of computer systems sufficiently simple that the small number of people who designed them completely understood every aspect of their operation was demonstrated on Apollo 14. As the Lunar Module was checked out prior to the landing, the astronauts in the spacecraft and Mission Control saw the abort signal come on, which was supposed to indicate the big Abort button on the control panel had been pushed. This button, if pressed during descent to the lunar surface, immediately aborted the landing attempt and initiated a return to lunar orbit. This was a “one and done” operation: no Microsoft-style “Do you really mean it?” tea ceremony before ending the mission. Tapping the switch made the signal come and go, and it was concluded the most likely cause was a piece of metal contamination floating around inside the switch and occasionally shorting the contacts. The abort signal caused no problems during lunar orbit, but if it should happen during descent, perhaps jostled by vibration from the descent engine, it would be disastrous: wrecking a mission costing hundreds of millions of dollars and, coming on the heels of Apollo 13's mission failure and narrow escape from disaster, possibly bring an end to the Apollo lunar landing programme.

The Lunar Module AGC team, with Don Eyles as the lead, was faced with an immediate challenge: was there a way to patch the software to ignore the abort switch, protecting the landing, while still allowing an abort to be commanded, if necessary, from the computer keyboard (DSKY)? The answer to this was obvious and immediately apparent: no. The landing software, like all AGC programs, ran from read-only rope memory which had been woven on the ground months before the mission and could not be changed in flight. But perhaps there was another way. Eyles and his colleagues dug into the program listing, traced the path through the logic, and cobbled together a procedure, then tested it in the simulator at the Instrumentation Laboratory. While the AGC's programming was fixed, the AGC operating system provided low-level commands which allowed the crew to examine and change bits in locations in the read-write memory. Eyles discovered that by setting the bit which indicated that an abort was already in progress, the abort switch would be ignored at the critical moments during the descent. As with all software hacks, this had other consequences requiring their own work-arounds, but by the time Apollo 14's Lunar Module emerged from behind the Moon on course for its landing, a complete procedure had been developed which was radioed up from Houston and worked perfectly, resulting in a flawless landing.

These and many other stories of the development and flight experience of the AGC lunar landing software are related here by the person who wrote most of it and supported every lunar landing mission as it happened. Where technical detail is required to understand what is happening, no punches are pulled, even to the level of bit-twiddling and hideously clever programming tricks such as using an overflow condition to skip over an EXTEND instruction, converting the following instruction from double precision to single precision, all in order to save around forty words of precious non-bank-switched memory. In addition, this is a personal story, set in the context of the turbulent 1960s and early ’70s, of the author and other young people accomplishing things no humans had ever before attempted.

It was a time when everybody was making it up as they went along, learning from experience, and improvising on the fly; a time when a person who had never written a line of computer code would write, as his first program, the code that would land men on the Moon, and when the creativity and hard work of individuals made all the difference. Already, by the end of the Apollo project, the curtain was ringing down on this era. Even though a number of improvements had been developed for the LM AGC software which improved precision landing capability, reduced the workload on the astronauts, and increased robustness, none of these were incorporated in the software for the final three Apollo missions, LUMINARY 210, which was deemed “good enough” and the benefit of the changes not worth the risk and effort to test and incorporate them. Programmers seeking this kind of adventure today will not find it at NASA or its contractors, but instead in the innovative “New Space” and smallsat industries.

November 2019 Permalink

Godwin, Robert ed. Gemini 6: The NASA Mission Reports. Burlington, Ontario, Canada: Apogee Books, 2000. ISBN 1-896522-61-0.

February 2001 Permalink

Godwin, Robert ed. Freedom 7: The NASA Mission Reports. Burlington, Ontario, Canada: Apogee Books, 2000. ISBN 1-896522-80-7.
This volume in the superb Apogee NASA Mission Reports series covers Alan Shepard's May 5th, 1961 suborbital flight in Freedom 7, the first U.S. manned space flight. Included are the press kit for the mission, complete transcripts of the post-flight debriefings and in-flight communications, and proceedings of a conference held in June 1961 to report mission results. In addition, the original 1958 request for astronaut volunteers (before it was decided that only military test pilots need apply) is reproduced, along with the press conference introducing the Mercury astronauts, which Tom Wolfe so vividly (and accurately) described in The Right Stuff. A bonus CD-ROM includes the complete in-flight films of the instrument panel and astronaut, a 30 minute NASA documentary about the flight, and the complete NASA official history of Project Mercury, This New Ocean, as a PDF document. There are few if any errors in the transcriptions of the documents. The caption for the photograph of Freedom 7 on the second page of colour plates makes the common error of describing its heat shield as “ablative fiberglass”. In fact, as stated on page 145, suborbital missions used a beryllium heat sink; only orbital capsules were equipped with the ablative shield.

December 2004 Permalink

Godwin, Robert ed. Friendship 7: The NASA Mission Reports. Burlington, Ontario, Canada: Apogee Books, 1999. ISBN 1-896522-60-2.
This installment in the Apogee NASA Mission Reports series contains original pre- and post-flight documents describing the first United States manned orbital flight piloted by John Glenn on February 20th, 1962, including a complete transcript of the air-to-ground communications from launch through splashdown. An excerpt from the Glenn's postflight debriefing describing his observations from space including the “fireflies” seen at orbital sunrise is included, along with a scientific evaluation which, in retrospect, seems to have gotten everything just about right. Glenn's own 13 page report on the flight is among the documents, as is backup pilot Scott Carpenter's report on training for the mission in which he describes the “extinctospectropolariscope-occulogyrogravoadaptometer”, abbreviated “V-Meter” in order to fit into the spacecraft (p. 110). A companion CD-ROM includes a one hour NASA film about the mission, with flight day footage from the tracking stations around the globe, and film from the pilot observation camera synchronised with recorded radio communications. An unintentionally funny introduction by the editor (complete with two idiot “it's”-es on consecutive lines) attempts to defend Glenn's 1998 political junket / P.R. stunt aboard socialist space ship Discovery. “If NASA is going to conduct gerontology experiments in orbit, who is more eminently qualified….” Well, a false predicate does imply anything, but if NASA were at all genuinely interested in geezers in space independent of political payback, why didn't they also fly John Young, only nine years Glenn's junior, who walked on the Moon, commanded the first flight of the space shuttle, was Chief of the Astronaut Office for ten years, and a NASA astronaut continuously from 1962 until his retirement in 2004, yet never given a flight assignment since 1983? Glenn's competence and courage needs no embellishment—and the contrast between the NASA in the days of his first flight and that of his second could not be more stark.

December 2005 Permalink

Guiteras, Daniel. Launch On Need. Unknown: T-Cell Books, 2010. ISBN 978-0-615-37221-1.
An almost universal convention of the alternative history genre is that there is a single point of departure (which I call “the veer”) where an event or fact in the narrative differs from that in the historical record, whence the rest of the story plays out with the same logic and plausibility as what actually happened in our timeline. When this is done well, it makes for engaging and thought-provoking fiction, as there are few things which so engage the cognitive veneer of our ancient brains as asking “what if?” This book is a superb exemplar of this genre, which works both as a thriller and an exploration of how the Space Shuttle program might have coped with the damage to orbiter Columbia due to foam shed from the bipod ramp of the external tank during its launch on STS-107.

Here, the veer is imagining NASA remained the kind of “can do”, “whatever it takes” organisation that it was in the early days of space flight through the rescue of Apollo 13 instead of the sclerotic bureaucracy it had become in the Shuttle era (and remains today). Dismissing evidence of damage to Columbia's thermal protection system (TPS) due to a foam strike, and not even seeking imagery from spy satellites, NASA's passive “managers” sighed and said “nothing could be done anyway” and allowed the crew to complete their mission and die during re-entry.

This needn't have happened. The Columbia Accident Investigation Board (CAIB) explored whether a rescue mission (PDF, scroll down to page 173), mounted as soon as possible after the possible damage to Columbia's TPS was detected, might have been able to rescue the crew before the expendables aboard Columbia were exhausted. Their conclusion? A rescue mission was possible, but only at the cost of cutting corners on safety margins and assuming nothing went wrong in the process of bringing the rescue shuttle, Atlantis, to the pad and launching her.

In this novel, the author takes great care to respect the dead, only referring to members of Columbia's crew by their crew positions such as “commander” or “mission specialist”, and invents names for those in NASA involved in the management of the actual mission. He draws upon the CAIB-envisioned rescue mission, including tables and graphics from their report, while humanising their dry prose with views of events as they unfold by fallible humans living them.

You knew this was coming, didn't you? You were waiting for it—confess! So here we go, into the quibbles. Some of these are substantial spoilers, so be warned.

Spoiler warning: Plot and/or ending details follow.  
Page numbers in the items below are from the Kindle edition, in which page numbers and their correspondence to print editions tend to be somewhat fluid. Consequently, depending upon how you arrive there, the page number in your edition may differ by ±1 page.

On p. 2, Brown “knew E208 was a high-resolution video camera…” which “By T-plus-240 seconds … had run through 1,000 feet of film.”
Video cameras do not use film. The confusion between video and film persists for several subsequent chapters.
On p. 5 the fifth Space Shuttle orbiter constructed is referred to as “Endeavor”.
In fact, this ship's name is properly spelled “Endeavour”, named after the Royal Navy research ship.
On p. 28 “…the crew members spent an additional 3,500 hundred hours studying and training…”
That's forty years—I think not.
On p. 55 Kalpana Chawla is described as a “female Indian astronaut.”
While Chawla was born in India, she became a U.S. citizen in 1990 and presumably relinquished her Indian citizenship in the process of naturalisation.
On p. 57 “Both [STS-107] astronauts selected for this EVA have previous spacewalk experience…”.
In fact, none of the STS-107 astronauts had ever performed an EVA.
On p. 65 “Normally, when spacewalks were part of the mission plan, the entire cabin of the orbiter was decompressed at least 24 hours prior to the start of the spacewalk.”
Are you crazy! EVA crewmembers pre-breathe pure oxygen in the cabin, then adapt to the low pressure of the spacesuit in the airlock, but the Shuttle cabin is never depressurised. If it were what would the other crewmembers breathe—Fireball XL5 oxygen pills?
On p. 75 the EVA astronaut looks out from Columbia's airlock and sees Cape Horn.
But the mission has been launched into an inclination of 39 degrees, so Cape Horn (55°59' S) should be out of sight to the South. Here is the view from Columbia's altitude on a pass over South America at the latitude of Cape Horn.
On p. 221 the countdown clock is said to have been “stuck on nine minutes zero seconds for the past three hours and twenty-seven minutes.”
The T−9 minute hold is never remotely that long. It's usually on the order of 10 to 20 minutes. If there were a reason for such a long hold, it would have been performed much earlier in the count. In any case, given the short launch window for the rendezvous, there'd be no reason for a long planned hold, and an unplanned hold would have resulted in a scrub of the mission until the next alignment with the plane of Columbia's orbit.
On p. 271 the crew of Atlantis open the payload bay doors shortly before the rendezvous with Columbia.
This makes no sense. Shuttles have to open their payload bay doors shortly after achieving orbit so that the radiators can discard heat. Atlantis would have opened its payload bay doors on the first orbit, not 24 hours later whilst approaching Columbia.
On p. 299 the consequences of blowing the crew ingress/egress hatch with the pyrotechnics is discussed.
There is no reason to consider doing this. From the inception of the shuttle program, the orbiter hatch has been able to be opened from the inside. The crew need only depressurise the orbiter and then operate the hatch opening mechanism.
On p. 332 “Standing by for communications blackout.”
The communications blackout is a staple of spaceflight drama but, in the shuttle era described in this novel, a thing of the past. While communications from the ground are blocked by plasma during reentry, communications from the shuttle routed through the TDRSS satellites are available throughout reentry except for brief periods when the orbiter's antennas are not aimed at the relay satellite overhead.
On p. 349 an Aegis guided missile cruiser shoots down the abandoned Columbia.
Where do I start? A space shuttle orbiter weighs about 100 tonnes. An SM-3 has a kinetic kill energy of around 130 megajoules, which is impressive, but is likely to pass through the structure of the shuttle, dispersing some debris, but leaving most of the mass behind. But let's suppose Columbia were dispersed into her component parts. Well, then the massive parts, such as the three main engines, would remain in orbit even longer, freed of the high-drag encumbrance of the rest of the structure, and come down hot and hard at random places around the globe. Probably, they'd splash in the ocean, but maybe they wouldn't—we'll never know.
Spoilers end here.  

While it's fun to spot and research goofs like these, I found they did not detract in any way from enjoyment of the novel, which is a perfectly plausible alternative history of Columbia's last mission.

February 2012 Permalink

Hall, Eldon C. Journey to the Moon: The History of the Apollo Guidance Computer. Reston, VA: AIAA, 1996. ISBN 1-56347-185-X.

September 2001 Permalink

Hall, R. Cargill. Lunar Impact. Washington: National Aeronautics and Space Administration, 1977. ISBN 978-0-486-47757-2. NASA SP-4210.
One of the wonderful things about the emergence of electronic books is that long out-of-print works from publishers' back-lists are becoming available once again since the cost of keeping them in print, after the initial conversion to an electronic format, is essentially zero. The U.S. civilian space agency NASA is to be commended for their efforts to make publications in their NASA history series available electronically at a bargain price. Many of these documents, chronicling the early days of space exploration from a perspective only a few years after the events, have been out of print for decades and some command forbidding prices on used book markets. Those interested in reading them, as opposed to collectors, now have an option as inexpensive as it is convenient to put these works in their hands.

The present volume, originally published in 1977, chronicles Project Ranger, NASA's first attempt to obtain “ground truth” about the surface of the Moon by sending probes to crash on its surface, radioing back high-resolution pictures, measuring its composition, and hard-landing scientific instruments on the surface to study the Moon's geology. When the project was begun in 1959, it was breathtakingly ambitious—so much so that one gets the sense those who set its goals did not fully appreciate the difficulty of accomplishing them. Ranger was to be not just a purpose-built lunar probe, but rather a general-purpose “bus” for lunar and planetary missions which could be equipped with different scientific instruments depending upon the destination and goals of the flight. It would incorporate, for the first time in a deep space mission, three-axis stabilisation, a steerable high-gain antenna, midcourse and terminal trajectory correction, an onboard (albeit extremely primitive) computer, real-time transmission of television imagery, support by a global Deep Space Network of tracking stations which did not exist before Ranger, sterilisation of the spacecraft to protect against contamination of celestial bodies by terrestrial organisms, and a retro-rocket and landing capsule which would allow rudimentary scientific instruments to survive thumping down on the Moon and transmit their results back to Earth.

This was a great deal to bite off, and as those charged with delivering upon these lofty goals discovered, extremely difficult to chew, especially in a period where NASA was still in the process of organising itself and lines of authority among NASA Headquarters, the Jet Propulsion Laboratory (charged with developing the spacecraft and conducting the missions) and the Air Force (which provided the Atlas-Agena launch vehicle that propelled Ranger to the Moon) were ill-defined and shifting frequently. This, along with the inherent difficulty of what was being attempted, contributed to results which can scarcely be imagined in an era of super-conservative mission design: six consecutive failures between 1961 and 1964, with a wide variety of causes. Even in the early days of spaceflight, this was enough to get the attention of the press, politicians, and public, and it was highly probable that had Ranger 7 also failed, it would be the end of the program. But it didn't—de-scoped to just a camera platform, it performed flawlessly and provided the first close-up glimpse of the Moon's surface. Rangers 8 and 9 followed, both complete successes, with the latter relaying pictures “live from the Moon” to televisions of viewers around the world. To this day I recall seeing them and experiencing a sense of wonder which is difficult to appreciate in our jaded age.

Project Ranger provided both the technology and experience base used in the Mariner missions to Venus, Mars, and Mercury. While the scientific results of Ranger were soon eclipsed by those of the Surveyor soft landers, it is unlikely that program would have succeeded without learning the painful lessons from Ranger.

The electronic edition of this book appears to have been created by scanning a print copy and running it through an optical character recognition program, then performing a spelling check and fixing errors it noted. However, no close proofreading appears to have been done, so that scanning errors which resulted in words in the spelling dictionary were not corrected. This results in a number of goofs in the text, some of which are humorous. My favourite is the phrase “midcourse correction bum [burn]” which occurs on several occasions. I imagine a dissipated wino with his trembling finger quivering above a big red “FIRE” button at a console at JPL. British readers may…no, I'm not going there. Illustrations from the original book are scanned and included as tiny thumbnails which cannot be enlarged. This is adequate for head shots of people, but for diagrams, charts, and photographs of hardware and the lunar surface, next to useless. References to endnotes in the text look like links but (at least reading the Kindle edition on an iPad) do nothing. These minor flaws do not seriously detract from the glimpse this work provides of unmanned planetary exploration at its moment of creation or the joy that this account is once again readily available.

Unlike many of the NASA history series, a paperback reprint edition is available, published by Dover. It is, however, much more expensive than the electronic edition.

Update: Reader J. Peterson writes that a free on-line edition of this book is available on NASA's Web site, in which the illustrations may be clicked to view full-resolution images.

February 2012 Permalink

Hendrickx, Bart and Bert Vis. Energiya-Buran. Chichester, UK: Springer Praxis, 2007. ISBN 978-0-387-69848-9.
This authoritative history chronicles one of the most bizarre episodes of the Cold War. When the U.S. Space Shuttle program was launched in 1972, the Soviets, unlike the majority of journalists and space advocates in the West who were bamboozled by NASA's propaganda, couldn't make any sense of the economic justification for the program. They worked the numbers, and they just didn't work—the flight rates, cost per mission, and most of the other numbers were obviously not achievable. So, did the Soviets chuckle at this latest folly of the capitalist, imperialist aggressors and continue on their own time-proven path of mass-produced low-technology expendable boosters? Well, of course not! They figured that even if their wisest double-domed analysts were unable to discern the justification for the massive expenditures NASA had budgeted for the Shuttle, there must be some covert military reason for its existence to which they hadn't yet twigged, and hence they couldn't tolerate a shuttle gap and consequently had to build their own, however pointless it looked on the surface.

And that's precisely what they did, as this book so thoroughly documents, with a detailed history, hundreds of pictures, and technical information which has only recently become available. Reasonable people can argue about the extent to which the Soviet shuttle was a copy of the American (and since the U.S. program started years before and placed much of its design data into the public domain, any wise designer would be foolish not to profit by using it), but what is not disputed is that (unlike the U.S. Shuttle) Energiya was a general purpose heavy-lift launcher which had the orbiter Buran as only one of its possible payloads and was one of the most magnificent engineering projects of the space programs of any nation, involving massive research and development, manufacturing, testing, integrated mission simulation, crew training, and flight testing programs.

Indeed, Energiya-Buran was in many ways a better-conceived program for space access than the U.S. Shuttle program: it integrated a heavy payload cargo launcher with the shuttle program, never envisioned replacing less costly expendable boosters with the shuttle, and forecast a development program which would encompass full reusability of boosters and core stages and both unmanned cargo and manned crew changeout missions to Soviet space stations.

The program came to a simultaneously triumphant and tragic end: the Energiya booster and the Energiya-Buran shuttle system performed flawless missions (the first Energiya launch failed to put its payload into orbit, but this was due to a software error in the payload: the launcher performed nominally from ignition through payload separation).

In the one and only flight of Buran (launch and landing video, other launch views) the orbiter was placed into its intended orbit and landed on the cosmodrome runway at precisely the expected time.

And then, in the best tradition not only of the Communist Party of the Soviet Union but of the British Labour Party of the 1970s, this singular success was rewarded by cancellation of the entire program. As an engineer, I have almost unlimited admiration for my ex-Soviet and Russian colleagues who did such masterful work and who will doubtless advance technology in the future to the benefit of us all. We should celebrate the achievement of those who created this magnificent space transportation system, while encouraging those inspired by it to open the high frontier to all of those who exulted in its success.

January 2009 Permalink

Hickam, Homer H., Jr. Rocket Boys. New York: Doubleday, 1998. ISBN 0-385-33321-8.
The author came of age in southern West Virginia during the dawn of the space age. Inspired by science fiction and the sight of Sputnik gliding through the patch of night sky between the mountains which surrounded his coal mining town, he and a group of close friends decided to build their own rockets. Counselled by the author's mother, “Don't blow yourself up”, they managed not only to avoid that downside of rocketry (although Mom's garden fence was not so lucky), but succeeded in building and launching more than thirty rockets powered by, as they progressed, first black powder, then melted saltpetre and sugar (“rocket candy”), and finally “zincoshine”, a mixture of powdered zinc and sulphur bound by 200 proof West Virginia mountain moonshine, which propelled their final rocket almost six miles into the sky. Their efforts won them the Gold and Silver award at the National Science Fair in 1960, and a ticket out of coal country for the author, who went on to a career as a NASA engineer. This is a memoir by a member of the last generation when the U.S. was still free enough for boys to be boys, and boys with dreams were encouraged to make them come true. This book will bring back fond memories for any member of that generation, and inspire envy among those who postdate that golden age.

This book served as the basis for the 1999 film October Sky, which I have not seen.

July 2005 Permalink

Hoagland, Richard C. and Mike Bara. Dark Mission. Los Angeles: Feral House, 2007. ISBN 1-932595-26-0.
Author Richard C. Hoagland first came to prominence as an “independent researcher” and advocate that “the face on Mars” was an artificially-constructed monument built by an ancient extraterrestrial civilisation. Hoagland has established himself as one of the most indefatigable and imaginative pseudoscientific crackpots on the contemporary scene, and this œuvre pulls it all together into a side-splittingly zany compendium of conspiracy theories, wacky physics, imaginative image interpretation, and feuds within the “anomalist” community—a tempest in a crackpot, if you like.

Hoagland seems to possess a visual system which endows him with a preternatural ability, undoubtedly valuable for an anomalist, of seeing things that aren't there. Now you may look at a print of a picture taken on the lunar surface by an astronaut with a Hasselblad camera and see, in the black lunar sky, negative scratches, film smudges, lens flare, and, in contrast-stretched and otherwise manipulated digitally scanned images, artefacts of the image processing filters applied, but Hoagland immediately perceives “multiple layers of breathtaking ‘structural construction’ embedded in the NASA frame; multiple surviving ‘cell-like rooms,’ three-dimensional ‘cross-bracing,’ angled ‘stringers,’ etc… all following logical structural patterns for a massive work of shattered, but once coherent, glass-like mega-engineering” (p. 153, emphasis in the original). You can see these wonders for yourself on Hoagland's site, The Enterprise Mission. From other Apollo images Hoagland has come to believe that much of the near side of the Moon is covered by the ruins of glass and titanium domes, some which still reach kilometres into the lunar sky and towered over some of the Apollo landing sites.

Now, you might ask, why did the Apollo astronauts not remark upon these prodigies, either while presumably dodging them when landing and flying back to orbit, nor on the surface, nor afterward. Well, you see, they must have been sworn to secrecy at the time and later (p. 176) hypnotised to cause them to forget the obvious evidence of a super-civilisation they were tripping over on the lunar surface. Yeah, that'll work.

Now, Occam's razor advises us not to unnecessarily multiply assumptions when formulating our hypotheses. On the one hand, we have the mainstream view that NASA missions have honestly reported the data they obtained to the public, and that these data, to date, include no evidence (apart from the ambiguous Viking biology tests on Mars) for extraterrestrial life nor artefacts of another civilisation. On the other, Hoagland argues:

  • NASA has been, from inception, ruled by three contending secret societies, all of which trace their roots to the gods of ancient Egypt: the Freemasons, unrepentant Nazi SS, and occult disciples of Aleister Crowley.
  • These cults have arranged key NASA mission events to occur at “ritual” times, locations, and celestial alignments. The Apollo 16 lunar landing was delayed due to a faked problem with the SPS engine so as to occur on Hitler's birthday.
  • John F. Kennedy was assassinated by a conspiracy including Lyndon Johnson and Congressman Albert Thomas of Texas because Kennedy was about to endorse a joint Moon mission with the Soviets, revealing to them the occult reasons behind the Apollo project.
  • There are two factions within NASA: the “owls”, who want to hide the evidence from the public, and the “roosters”, who are trying to get it out by covert data releases and cleverly coded clues.

    But wait, there's more!

  • The energy of the Sun comes, at least in part, from a “hyperdimensional plane” which couples to rotating objects through gravitational torsion (you knew that was going to come in sooner or later!) This energy expresses itself through a tetrahedral geometry, and explains, among other mysteries, the Great Red Spot of Jupiter, the Great Dark Spot of Neptune, Olympus Mons on Mars, Mauna Kea in Hawaii, and the precession of isolated pulsars.
  • The secrets of this hyperdimensional physics, glimpsed by James Clerk Maxwell in his quaternion (check off another crackpot checklist item) formulation of classical electrodynamics, were found by Hoagland to be encoded in the geometry of the “monuments” of Cydonia on Mars.
  • Mars was once the moon of a “Planet V”, which exploded (p. 362).

    And that's not all!

  • NASA's Mars rover Opportunity imaged a fossil in a Martian rock and then promptly ground it to dust.
  • The terrain surrounding the rover Spirit is littered with artificial objects.
  • Mars Pathfinder imaged a Sphinx on Mars.

    And if that weren't enough!

  • Apollo 17 astronauts photographed the head of an anthropomorphic robot resembling C-3PO lying in Shorty Crater on the Moon (p. 487).

It's like Velikovsky meets The Illuminatus! Trilogy, with some of the darker themes of “Millennium” thrown in for good measure.

Now, I'm sure, as always happens when I post a review like this, the usual suspects are going to write to demand whatever possessed me to read something like this and/or berate me for giving publicity to such hyperdimensional hogwash. Lighten up! I read for enjoyment, and to anybody with a grounding in the Actual Universe™, this stuff is absolutely hilarious: there's a chortle every few pages and a hearty guffaw or two in each chapter. The authors actually write quite well: this is not your usual semi-literate crank-case sludge, although like many on the far fringes of rationality they seem to be unduly challenged by the humble apostrophe. Hoagland is inordinately fond of the word “infamous”, but this becomes rather charming after the first hundred or so, kind of like the verbal tics of your crazy uncle, who Hoagland rather resembles. It's particularly amusing to read the accounts of Hoagland's assorted fallings out and feuds with other “anomalists”; when Tom Van Flandern concludes you're a kook, then you know you're out there, and I don't mean hanging with the truth.

December 2007 Permalink

Jenkins, Dennis R. and Jorge R. Frank. The Apollo 11 Moon Landing. North Branch, MN: Specialty Press, 2009. ISBN 978-1-58007-148-2.
This book, issued to commemorate the 40th anniversary of the Apollo 11 Moon landing, is a gorgeous collection of photographs, including a number of panoramas digitally assembled from photos taken during the mission which appear here for the first time. The images cover all aspects of the mission: the evolution of the Apollo project, crew training, stacking the launcher and spacecraft, voyage to the Moon, surface operations, and return to Earth. The photos have accurate and informative captions, and each chapter includes a concise but comprehensive description of its topic.

This is largely a picture book, and almost entirely focused upon the Apollo 11 mission, not the Apollo program as a whole. Unless you are an absolute space nut (guilty as charged), you will almost certainly see pictures here you've never seen before, including Neil Armstrong's brush with death when the Lunar Landing Research Vehicle went all pear shaped and he had to punch out (p. 35). Look at how the ejection seat motor vectored to buy him altitude for the chute to open!

Did you know that the iconic image of Buzz Aldrin on the Moon was retouched (or, as we'd say today, PhotoShopped)? No, I'm not talking about a Moon hoax, but just that Neil Armstrong, with his Hasselblad camera and no viewfinder, did what so many photographers do—he cut off Aldrin's head in the picture. NASA public affairs folks “reconstructed” the photo that Armstrong meant to take, but whilst airbrushing the top of the helmet, they forgot to include the OPS VHF antenna which extends from Aldrin's backpack in many other photos taken on the lunar surface.

This is a great book, and a worthy commemoration of the achievement of Apollo 11. It, of course, only scratches the surface of the history of the Apollo program, or even the details of Apollo 11 mission, but I don't know an another source which brings together so many images which evoke that singular exploit. The Introduction includes a list of sources for further reading which I was amazed (or maybe not) to discover that all of which I had read.

August 2009 Permalink

Kaufman, Marc. First Contact. New York: Simon & Schuster, 2011. ISBN 978-1-4391-0901-4.
How many fields of science can you think of which study something for which there is no generally accepted experimental evidence whatsoever? Such areas of inquiry certainly exist: string theory and quantum gravity come immediately to mind, but those are research programs motivated by self-evident shortcomings in the theoretical foundations of physics which become apparent when our current understanding is extrapolated to very high energies. Astrobiology, the study of life in the cosmos, has, to date, only one exemplar to investigate: life on Earth. For despite the enormous diversity of terrestrial life, it shares a common genetic code and molecular machinery, and appears to be descended from a common ancestral organism.

And yet in the last few decades astrobiology has been a field which, although having not so far unambiguously identified extraterrestrial life, has learned a great deal about life on Earth, the nature of life, possible paths for the origin of life on Earth and elsewhere, and the habitats in the universe where life might be found. This book, by a veteran Washington Post science reporter, visits the astrobiologists in their native habitats, ranging from deep mines in South Africa, where organisms separated from the surface biosphere for millions of years have been identified, Antarctica; whose ice hosts microbes the likes of which might flourish on the icy bodies of the outer solar system; to planet hunters patiently observing stars from the ground and space to discover worlds orbiting distant stars.

It is amazing how much we have learned in such a short time. When I was a kid, many imagined that Venus's clouds shrouded a world of steamy jungles, and that Mars had plants which changed colour with the seasons. No planet of another star had been detected, and respectable astronomers argued that the solar system might have been formed by a freak close approach between two stars and that planets might be extremely rare. The genetic code of life had not been decoded, and an entire domain of Earthly life, bearing important clues for life's origin, was unknown and unsuspected. This book describes the discoveries which have filled in the blanks over the last few decades, painting a picture of a galaxy in which planets abound, many in the “habitable zone” of their stars. Life on Earth has been found to have colonised habitats previously considered as inhospitable to life as other worlds: absence of oxygen, no sunlight, temperatures near freezing or above the boiling point of water, extreme acidity or alkalinity: life finds a way.

We may have already discovered extraterrestrial life. The author meets the thoroughly respectable scientists who operated the life detection experiments of the Viking Mars landers in the 1970s, sought microfossils of organisms in a meteorite from Mars found in Antarctica, and searched for evidence of life in carbonaceous meteorites. Each believes the results of their work is evidence of life beyond Earth, but the standard of evidence required for such an extraordinary claim has not been met in the opinion of most investigators.

While most astrobiologists seek evidence of simple life forms (which exclusively inhabited Earth for most of its history), the Search for Extraterrestrial Intelligence (SETI) jumps to the other end of evolution and seeks interstellar communications from other technological civilisations. While initial searches were extremely limited in the assumptions about signals they might detect, progress in computing has drastically increased the scope of these investigations. In addition, other channels of communication, such as very short optical pulses, are now being explored. While no signals have been detected in 50 years of off and on searching, only a minuscule fraction of the search space has been explored, and it may be that in retrospect we'll realise that we've had evidence of interstellar signals in our databases for years in the form of transient pulses not recognised because we were looking for narrowband continuous beacons.

Discovery of life beyond the Earth, whether humble microbes on other bodies of the solar system or an extraterrestrial civilisation millions of years older than our own spamming the galaxy with its ETwitter feed, would arguably be the most significant discovery in the history of science. If we have only one example of life in the universe, its origin may have been a forbiddingly improbable fluke which happened only once in our galaxy or in the entire universe. But if there are two independent examples of the origin of life (note that if we find life on Mars, it is crucial to determine whether it shares a common origin with terrestrial life: since meteors exchange material between the planets, it's possible Earth life originated on Mars or vice versa), then there is every reason to believe life is as common in the cosmos as we are now finding planets to be. Perhaps in the next few decades we will discover the universe to be filled with wondrous creatures awaiting our discovery. Or maybe not—we may be alone in the universe, in which case it is our destiny to bring it to life.

November 2013 Permalink

Kelly, Thomas J. Moon Lander. Washington: Smithsonian Institution Press, 2001. ISBN 1-56098-998-X.

January 2003 Permalink

Kennedy, Gregory P. The Rockets and Missiles of White Sands Proving Ground, 1945–1958. Atglen, PA: Schiffer Military History, 2009. ISBN 978-0-7643-3251-7.
Southern New Mexico has been a centre of American rocketry from its origin to the present day. After being chased out of Massachusetts due to his inventions' proclivity for making ear-shattering detonations and starting fires, Robert Goddard moved his liquid fuel rocket research to a site near Roswell, New Mexico in 1930 and continued to launch increasingly advanced rockets from that site until 1943, when he left to do war work for the Navy. Faced with the need for a range to test the missiles developed during World War II, in February 1945 the U.S. Army acquired a site stretching 100 miles north from the Texas-New Mexico border near El Paso and 41 miles east-west at the widest point, designated the “White Sands Proving Ground”: taking its name from the gypsum sands found in the region, also home to the White Sands National Monument.

Although established before the end of the war to test U.S. missiles, the first large rockets launched at the site were captured German V-2s (December 2002), with the first launched (unsuccessfully) in April 1946. Over the next six years, around seventy V-2s lifted off from White Sands, using the V-2's massive (for the time) one ton payload capacity to carry a wide variety of scientific instruments into the upper atmosphere and the edge of space. In the Bumper project, the V-2 was used as the booster for the world's first two stage liquid rocket, with its WAC Corporal second stage attaining an altitude of 248 miles: higher than some satellites orbit today (it did not, of course, attain anything near orbital velocity, and quickly fell back to Earth).

Simultaneously with launches of the V-2, U.S. rocketeers arrived at White Sands to test their designs—almost every U.S. missile of the 1940s and 1950s made its first flight there. These included research rockets such as Viking and Aerobee (first launched in 1948, it remained in service until 1985 with a total of 1037 launched); the Corporal, Sergeant, and Redstone ballistic missiles; Loki, Nike, Hawk anti-aircraft missiles; and a variety of tactical missiles including the unguided (!) nuclear-tipped Honest John.

White Sands in the forties and fifties was truly the Wild West of rocketry. Even by the standards of fighter aircraft development in the epoch, this was by guess and by gosh engineering in its purest incarnation. Consider Viking 8, which broke loose from the launch pad during a static test when hold-down fittings failed, and was allowed to fly to 20,000 feet to see what would happen (p. 97). Or Viking 10, whose engine exploded on the launch pad and then threatened a massive explosion because leaking fuel was causing the tankage to crumple as it left a vacuum. An intrepid rocketeer was sent out of the blockhouse with a carbine to shoot a hole in the top of the fuel tank and allow air to enter (p. 100)—problem solved! (The rocket was rebuilt and later flew successfully.) Then there was the time they ran out of 90% hydrogen peroxide and were told the first Viking launch would have to be delayed for two weeks until a new shipment could arrive by rail. Can't have that! So two engineers drove a drum of the highly volatile and corrosive substance in the back of a station wagon from Buffalo, New York to White Sands to meet the launch deadline (p. 79). In the Nike program, people worried about whether its aniline fuel would be sufficiently available under tactical conditions, so they tried using gasoline as fuel instead—BOOM! Nope, guess not (p. 132). With all this “innovation” going on, they needed a suitable place from which to observe it, so the pyramid-shaped blockhouse had reinforced concrete walls ten feet thick with a roof 27 feet thick at the peak. This was designed to withstand a direct impact from a V-2 falling from an altitude of 100 miles. “Once the rockets are up, who cares where they come down?”

And the pace of rockets going up was absolutely frenetic, almost inconceivable by the standards of today's hangar queens and launch pad prima donnas (some years ago, a booster which sat on the pad for more than a year was nicknamed the “civil servant”: it won't work and you can't fire it). By contrast, a single development program, the Loki anti-aircraft missile, conducted a total of 2282 launches at White Sands in 1953 and 1954 (p. 115)—that's an average of more than three a day, counting weekends and holidays!

The book concludes in 1958 when White Sands Proving Ground became White Sands Missile Range (scary pop-up at this link), which remains a centre of rocket development and testing to this day. With the advent of NASA and massively funded, long-term military procurement programs, much of the cut, try, and run like Hell days of rocketry came to a close; this book covers that period which, if not a golden age, was a heck of a lot of fun for engineers who enjoy making loud noises and punching holes in the sky.

The book is gorgeous, printed on glossy paper, with hundreds of illustrations. I noted no typographical or factual errors. A complete list of all U.S. V-2, WAC Corporal, and Viking launches is given in appendices at the end.

May 2010 Permalink

Kluger, Jeffrey. Apollo 8. New York: Picador, 2017. ISBN 978-1-250-18251-7.
As the tumultuous year 1968 drew to a close, NASA faced a serious problem with the Apollo project. The Apollo missions had been carefully planned to test the Saturn V booster rocket and spacecraft (Command/Service Module [CSM] and Lunar Module [LM]) in a series of increasingly ambitious missions, first in low Earth orbit (where an immediate return to Earth was possible in case of problems), then in an elliptical Earth orbit which would exercise the on-board guidance and navigation systems, followed by lunar orbit, and finally proceeding to the first manned lunar landing. The Saturn V had been tested in two unmanned “A” missions: Apollo 4 in November 1967 and Apollo 6 in April 1968. Apollo 5 was a “B” mission, launched on a smaller Saturn 1B booster in January 1968, to test an unmanned early model of the Lunar Module in low Earth orbit, primarily to verify the operation of its engines and separation of the descent and ascent stages. Apollo 7, launched in October 1968 on a Saturn 1B, was the first manned flight of the Command and Service modules and tested them in low Earth orbit for almost 11 days in a “C” mission.

Apollo 8 was planned to be the “D” mission, in which the Saturn V, in its first manned flight, would launch the Command/Service and Lunar modules into low Earth orbit, where the crew, commanded by Gemini veteran James McDivitt, would simulate the maneuvers of a lunar landing mission closer to home. McDivitt's crew was trained and ready to go in December 1968. Unfortunately, the lunar module wasn't. The lunar module scheduled for Apollo 8, LM-3, had been delivered to the Kennedy Space Center in June of 1968, but was, to put things mildly, a mess. Testing at the Cape discovered more than a hundred serious defects, and by August it was clear that there was no way LM-3 would be ready for a flight in 1968. In fact, it would probably slip to February or March 1969. This, in turn, would push the planned “E” mission, for which the crew of commander Frank Borman, command module pilot James Lovell, and lunar module pilot William Anders were training, aimed at testing the Command/Service and Lunar modules in an elliptical Earth orbit venturing as far as 7400 km from the planet and originally planned for March 1969, three months later, to June, delaying all subsequent planned missions and placing the goal of landing before the end of 1969 at risk.

But NASA were not just racing the clock—they were also racing the Soviet Union. Unlike Apollo, the Soviet space program was highly secretive and NASA had to go on whatever scraps of information they could glean from Soviet publications, the intelligence community, and independent tracking of Soviet launches and spacecraft in flight. There were, in fact, two Soviet manned lunar programmes running in parallel. The first, internally called the Soyuz 7K-L1 but dubbed “Zond” for public consumption, used a modified version of the Soyuz spacecraft launched on a Proton booster and was intended to carry two cosmonauts on a fly-by mission around the Moon. The craft would fly out to the Moon, use its gravity to swing around the far side, and return to Earth. The Zond lacked the propulsion capability to enter lunar orbit. Still, success would allow the Soviets to claim the milestone of first manned mission to the Moon. In September 1968 Zond 5 successfully followed this mission profile and safely returned a crew cabin containing tortoises, mealworms, flies, and plants to Earth after their loop around the Moon. A U.S. Navy destroyer observed recovery of the re-entry capsule in the Indian Ocean. Clearly, this was preparation for a manned mission which might occur on any lunar launch window.

(The Soviet manned lunar landing project was actually far behind Apollo, and would not launch its N1 booster on that first, disastrous, test flight until February 1969. But NASA did not know this in 1968.) Every slip in the Apollo program increased the probability of its being scooped so close to the finish line by a successful Zond flyby mission.

These were the circumstances in August 1968 when what amounted to a cabal of senior NASA managers including George Low, Chris Kraft, Bob Gilruth, and later joined by Wernher von Braun and chief astronaut Deke Slayton, began working on an alternative. They plotted in secret, beneath the radar and unbeknownst to NASA administrator Jim Webb and his deputy for manned space flight, George Mueller, who were both out of the country, attending an international conference in Vienna. What they were proposing was breathtaking in its ambition and risk. They envisioned taking Frank Borman's crew, originally scheduled for Apollo 9, and putting them into an accelerated training program to launch on the Saturn V and Apollo spacecraft currently scheduled for Apollo 8. They would launch without a Lunar Module, and hence be unable to land on the Moon or test that spacecraft. The original idea was to perform a Zond-like flyby, but this was quickly revised to include going into orbit around the Moon, just as a landing mission would do. This would allow retiring the risk of many aspects of the full landing mission much earlier in the program than originally scheduled, and would also allow collection of precision data on the lunar gravitational field and high resolution photography of candidate landing sites to aid in planning subsequent missions. The lunar orbital mission would accomplish all the goals of the originally planned “E” mission and more, allowing that mission to be cancelled and therefore not requiring an additional booster and spacecraft.

But could it be done? There were a multitude of requirements, all daunting. Borman's crew, training toward a launch in early 1969 on an Earth orbit mission, would have to complete training for the first lunar mission in just sixteen weeks. The Saturn V booster, which suffered multiple near-catastrophic engine failures in its second flight on Apollo 6, would have to be cleared for its first manned flight. Software for the on-board guidance computer and for Mission Control would have to be written, tested, debugged, and certified for a lunar mission many months earlier than previously scheduled. A flight plan for the lunar orbital mission would have to be written from scratch and then tested and trained in simulations with Mission Control and the astronauts in the loop. The decision to fly Borman's crew instead of McDivitt's was to avoid wasting the extensive training the latter crew had undergone in LM systems and operations by assigning them to a mission without an LM. McDivitt concurred with this choice: while it might be nice to be among the first humans to see the far side of the Moon with his own eyes, for a test pilot the highest responsibility and honour is to command the first flight of a new vehicle (the LM), and he would rather skip the Moon mission and fly later than lose that opportunity. If the plan were approved, Apollo 8 would become the lunar orbit mission and the Earth orbit test of the LM would be re-designated Apollo 9 and fly whenever the LM was ready.

While a successful lunar orbital mission on Apollo 8 would demonstrate many aspects of a full lunar landing mission, it would also involve formidable risks. The Saturn V, making only its third flight, was coming off a very bad outing in Apollo 6 whose failures might have injured the crew, damaged the spacecraft hardware, and precluded a successful mission to the Moon. While fixes for each of these problems had been implemented, they had never been tested in flight, and there was always the possibility of new problems not previously seen.

The Apollo Command and Service modules, which would take them to the Moon, had not yet flown a manned mission and would not until Apollo 7, scheduled for October 1968. Even if Apollo 7 were a complete success (which was considered a prerequisite for proceeding), Apollo 8 would be only the second manned flight of the Apollo spacecraft, and the crew would have to rely upon the functioning of its power generation, propulsion, and life support systems for a mission lasting six days. Unlike an Earth orbit mission, if something goes wrong en route to or returning from the Moon, you can't just come home immediately. The Service Propulsion System on the Service Module would have to work perfectly when leaving lunar orbit or the crew would be marooned forever or crash on the Moon. It would only have been tested previously in one manned mission and there was no backup (although the single engine did incorporate substantial redundancy in its design).

The spacecraft guidance, navigation, and control system and its Apollo Guidance Computer hardware and software, upon which the crew would have to rely to navigate to and from the Moon, including the critical engine burns to enter and leave lunar orbit while behind the Moon and out of touch with Mission Control, had never been tested beyond Earth orbit.

The mission would go to the Moon without a Lunar Module. If a problem developed en route to the Moon which disabled the Service Module (as would happen to Apollo 13 in April 1970), there would be no LM to serve as a lifeboat and the crew would be doomed.

When the high-ranking conspirators presented their audacious plan to their bosses, the reaction was immediate. Manned spaceflight chief Mueller immediately said, “Can't do that! That's craziness!” His boss, administrator James Webb, said “You try to change the entire direction of the program while I'm out of the country?” Mutiny is a strong word, but this seemed to verge upon it. Still, Webb and Mueller agreed to meet with the lunar cabal in Houston on August 22. After a contentious meeting, Webb agreed to proceed with the plan and to present it to President Johnson, who was almost certain to approve it, having great confidence in Webb's management of NASA. The mission was on.

It was only then that Borman and his crewmembers Lovell and Anders learned of their reassignment. While Anders was disappointed at the prospect of being the Lunar Module Pilot on a mission with no Lunar Module, the prospect of being on the first flight to the Moon and entrusted with observation and photography of lunar landing sites more than made up for it. They plunged into an accelerated training program to get ready for the mission.

NASA approached the mission with its usual “can-do” approach and public confidence, but everybody involved was acutely aware of the risks that were being taken. Susan Borman, Frank's wife, privately asked Chris Kraft, director of Flight Operations and part of the group who advocated sending Apollo 8 to the Moon, with a reputation as a plain-talking straight shooter, “I really want to know what you think their chances are of coming home.” Kraft responded, “You really mean that, don't you?” “Yes,” she replied, “and you know I do.” Kraft answered, “Okay. How's fifty-fifty?” Those within the circle, including the crew, knew what they were biting off.

The launch was scheduled for December 21, 1968. Everybody would be working through Christmas, including the twelve ships and thousands of sailors in the recovery fleet, but lunar launch windows are set by the constraints of celestial mechanics, not human holidays. In November, the Soviets had flown Zond 6, and it had demonstrated the “double dip” re-entry trajectory required for human lunar missions. There were two system failures which killed the animal test subjects on board, but these were covered up and the mission heralded as a great success. From what NASA knew, it was entirely possible the next launch would be with cosmonauts bound for the Moon.

Space launches were exceptional public events in the 1960s, and the first flight of men to the Moon, just about a hundred years after Jules Verne envisioned three men setting out for the Moon from central Florida in a “cylindro-conical projectile” in De la terre à la lune (From the Earth to the Moon), similarly engaging the world, the launch of Apollo 8 attracted around a quarter of a million people to watch the spectacle in person and hundreds of millions watching on television both in North America and around the globe, thanks to the newfangled technology of communication satellites. Let's tune in to CBS television and relive this singular event with Walter Cronkite.

CBS coverage of the Apollo 8 launch

Now we step inside Mission Control and listen in on the Flight Director's audio loop during the launch, illustrated with imagery and simulations.

The Saturn V performed almost flawlessly. During the second stage burn mild pogo oscillations began but, rather than progressing to the point where they almost tore the rocket apart as had happened on the previous Saturn V launch, von Braun's team's fixes kicked in and seconds later Borman reported, “Pogo's damping out.” A few minutes later Apollo 8 was in Earth orbit.

Jim Lovell had sixteen days of spaceflight experience across two Gemini missions, one of them Gemini 7 where he endured almost two weeks in orbit with Frank Borman. Bill Anders was a rookie, on his first space flight. Now weightless, all three were experiencing a spacecraft nothing like the cramped Mercury and Gemini capsules which you put on as much as boarded. The Apollo command module had an interior volume of six cubic metres (218 cubic feet, in the quaint way NASA reckons things) which may not seem like much for a crew of three, but in weightlessness, with every bit of space accessible and usable, felt quite roomy. There were five real windows, not the tiny portholes of Gemini, and plenty of space to move from one to another.

With all this roominess and mobility came potential hazards, some verging on slapstick, but, in space, serious nonetheless. NASA safety personnel had required the astronauts to wear life vests over their space suits during the launch just in case the Saturn V malfunctioned and they ended up in the ocean. While moving around the cabin to get to the navigation station after reaching orbit, Lovell, who like the others hadn't yet removed his life vest, snagged its activation tab on a strut within the cabin and it instantly inflated. Lovell looked ridiculous and the situation comical, but it was no laughing matter. The life vests were inflated with carbon dioxide which, if released in the cabin, would pollute their breathing air and removal would use up part of a CO₂ scrubber cartridge, of which they had a limited supply on board. Lovell finally figured out what to do. After being helped out of the vest, he took it down to the urine dump station in the lower equipment bay and vented it into a reservoir which could be dumped out into space. One problem solved, but in space you never know what the next surprise might be.

The astronauts wouldn't have much time to admire the Earth through those big windows. Over Australia, just short of three hours after launch, they would re-light the engine on the third stage of the Saturn V for the “trans-lunar injection” (TLI) burn of 318 seconds, which would accelerate the spacecraft to just slightly less than escape velocity, raising its apogee so it would be captured by the Moon's gravity. After housekeeping (presumably including the rest of the crew taking off those pesky life jackets, since there weren't any wet oceans where they were going) and reconfiguring the spacecraft and its computer for the maneuver, they got the call from Houston, “You are go for TLI.” They were bound for the Moon.

The third stage, which had failed to re-light on its last outing, worked as advertised this time, with a flawless burn. Its job was done; from here on the astronauts and spacecraft were on their own. The booster had placed them on a free-return trajectory. If they did nothing (apart from minor “trajectory correction maneuvers” easily accomplished by the spacecraft's thrusters) they would fly out to the Moon, swing around its far side, and use its gravity to slingshot back to the Earth (as Lovell would do two years later when he commanded Apollo 13, although there the crew had to use the engine of the LM to get back onto a free-return trajectory after the accident).

Apollo 8 rapidly climbed out of the Earth's gravity well, trading speed for altitude, and before long the astronauts beheld a spectacle no human eyes had glimpsed before: an entire hemisphere of Earth at once, floating in the inky black void. On board, there were other concerns: Frank Borman was puking his guts out and having difficulties with the other end of the tubing as well. Borman had logged more than six thousand flight hours in his career as a fighter and test pilot, most of it in high-performance jet aircraft, and fourteen days in space on Gemini 7 without any motion sickness. Many people feel queasy when they experience weightlessness the first time, but this was something entirely different and new in the American space program. And it was very worrisome. The astronauts discussed the problem on private tapes they could downlink to Mission Control without broadcasting to the public, and when NASA got around to playing the tapes, the chief flight surgeon, Dr. Charles Berry, became alarmed.

As he saw it, there were three possibilities: motion sickness, a virus of some kind, or radiation sickness. On its way to the Moon, Apollo 8 passed directly through the Van Allen radiation belts, spending two hours in this high radiation environment, the first humans to do so. The total radiation dose was estimated as roughly the same as one would receive from a chest X-ray, but the composition of the radiation was different and the exposure was over an extended time, so nobody could be sure it was safe. The fact that Lovell and Anders had experienced no symptoms argued against the radiation explanation. Berry concluded that a virus was the most probable cause and, based upon the mission rules said, “I'm recommending that we consider canceling the mission.” The risk of proceeding with the commander unable to keep food down and possibly carrying a virus which the other astronauts might contract was too great in his opinion. This recommendation was passed up to the crew. Borman, usually calm and collected even by astronaut standards, exclaimed, “What? That is pure, unadulterated horseshit.” The mission would proceed, and within a day his stomach had settled.

This was the first case of space adaptation syndrome to afflict an American astronaut. (Apparently some Soviet cosmonauts had been affected, but this was covered up to preserve their image as invincible exemplars of the New Soviet Man.) It is now known to affect around a third of people experiencing weightlessness in environments large enough to move around, and spontaneously clears up in two to four (miserable) days.

The two most dramatic and critical events in Apollo 8's voyage would occur on the far side of the Moon, with 3500 km of rock between the spacecraft and the Earth totally cutting off all communications. The crew would be on their own, aided by the computer and guidance system and calculations performed on the Earth and sent up before passing behind the Moon. The first would be lunar orbit insertion (LOI), scheduled for 69 hours and 8 minutes after launch. The big Service Propulsion System (SPS) engine (it was so big—twice as large as required for Apollo missions as flown—because it was designed to be able to launch the entire Apollo spacecraft from the Moon if a “direct ascent” mission mode had been selected) would burn for exactly four minutes and seven seconds to bend the spacecraft's trajectory around the Moon into a closed orbit around that world.

If the SPS failed to fire for the LOI burn, it would be a huge disappointment but survivable. Apollo 8 would simply continue on its free-return trajectory, swing around the Moon, and fall back to Earth where it would perform a normal re-entry and splashdown. But if the engine fired and cut off too soon, the spacecraft would be placed into an orbit which would not return them to Earth, marooning the crew in space to die when their supplies ran out. If it burned just a little too long, the spacecraft's trajectory would intersect the surface of the Moon—lithobraking is no way to land on the Moon.

When the SPS engine shut down precisely on time and the computer confirmed the velocity change of the burn and orbital parameters, the three astronauts were elated, but they were the only people in the solar system aware of the success. Apollo 8 was still behind the Moon, cut off from communications. The first clue Mission Control would have of the success or failure of the burn would be when Apollo 8's telemetry signal was reacquired as it swung around the limb of the Moon. If too early, it meant the burn had failed and the spacecraft was coming back to Earth; that moment passed with no signal. Now tension mounted as the clock ticked off the seconds to the time expected for a successful burn. If that time came and went with no word from Apollo 8, it would be a really bad day. Just on time, the telemetry signal locked up and Jim Lovell reported, “Go ahead, Houston, this is Apollo 8. Burn complete. Our orbit 160.9 by 60.5.” (Lovell was using NASA's preferred measure of nautical miles; in proper units it was 311 by 112 km. The orbit would subsequently be circularised by another SPS burn to 112.7 by 114.7 km.) The Mission Control room erupted into an un-NASA-like pandemonium of cheering.

Apollo 8 would orbit the Moon ten times, spending twenty hours in a retrograde orbit with an inclination of 12 degrees to the lunar equator, which would allow it to perform high-resolution photography of candidate sites for early landing missions under lighting conditions similar to those expected at the time of landing. In addition, precision tracking of the spacecraft's trajectory in lunar orbit would allow mapping of the Moon's gravitational field, including the “mascons” which perturb the orbits of objects in low lunar orbits and would be important for longer duration Apollo orbital missions in the future.

During the mission, the crew were treated to amazing sights and, in particular, the dramatic difference between the near side, with its many flat “seas”, and the rugged highlands of the far side. Coming around the Moon they saw the spectacle of earthrise for the first time and, hastily grabbing a magazine of colour film and setting aside the planned photography schedule, Bill Anders snapped the photo of the Earth rising above the lunar horizon which became one of the most iconic photographs of the twentieth century. Here is a reconstruction of the moment that photo was taken.

On the ninth and next-to-last orbit, the crew conducted a second television transmission which was broadcast worldwide. It was Christmas Eve on much of the Earth, and, coming at the end of the chaotic, turbulent, and often tragic year of 1968, it was a magical event, remembered fondly by almost everybody who witnessed it and felt pride for what the human species had just accomplished.

You have probably heard this broadcast from the Moon, often with the audio overlaid on imagery of the Moon from later missions, with much higher resolution than was actually seen in that broadcast. Here, in three parts, is what people, including this scrivener, actually saw on their televisions that enchanted night. The famous reading from Genesis is in the third part. This description is eerily similar to that in Jules Verne's 1870 Autour de la lune.

After the end of the broadcast, it was time to prepare for the next and absolutely crucial maneuver, also performed on the far side of the Moon: trans-Earth injection, or TEI. This would boost the spacecraft out of lunar orbit and send it back on a trajectory to Earth. This time the SPS engine had to work, and perfectly. If it failed to fire, the crew would be trapped in orbit around the Moon with no hope of rescue. If it cut off too soon or burned too long, or the spacecraft was pointed in the wrong direction when it fired, Apollo 8 would miss the Earth and orbit forever far from its home planet or come in too steep and burn up when it hit the atmosphere. Once again the tension rose to a high pitch in Mission Control as the clock counted down to the two fateful times: this time they'd hear from the spacecraft earlier if it was on its way home and later or not at all if things had gone tragically awry. Exactly when expected, the telemetry screens came to life and a second later Jim Lovell called, “Houston, Apollo 8. Please be informed there is a Santa Claus.”

Now it was just a matter of falling the 375,000 kilometres from the Moon, hitting the precise re-entry corridor in the Earth's atmosphere, executing the intricate “double dip” re-entry trajectory, and splashing down near the aircraft carrier which would retrieve the Command Module and crew. Earlier unmanned tests gave confidence it would all work, but this was the first time men would be trying it.

There was some unexpected and embarrassing excitement on the way home. Mission Control had called up a new set of co-ordinates for the “barbecue roll” which the spacecraft executed to even out temperature. Lovell was asked to enter “verb 3723, noun 501” into the computer. But, weary and short on sleep, he fat-fingered the commands and entered “verb 37, noun 01”. This told the computer the spacecraft was back on the launch pad, pointing straight up, and it immediately slewed to what it thought was that orientation. Lovell quickly figured out what he'd done, “It was my goof”, but by this time he'd “lost the platform”: the stable reference the guidance system used to determine in which direction the spacecraft was pointing in space. He had to perform a manual alignment, taking sightings on a number of stars, to recover the correct orientation of the stable platform. This was completely unplanned but, as it happens, in doing so Lovell acquired experience that would prove valuable when he had to perform the same operation in much more dire circumstances on Apollo 13 after an explosion disabled the computer and guidance system in the Command Module. Here is the author of the book, Jeffrey Kluger, discussing Jim Lovell's goof.

The re-entry went completely as planned, flown entirely under computer control, with the spacecraft splashing into the Pacific Ocean just 6 km from the aircraft carrier Yorktown. But because the splashdown occurred before dawn, it was decided to wait until the sky brightened to recover the crew and spacecraft. Forty-three minutes after splashdown, divers from the Yorktown arrived at the scene, and forty-five minutes after that the crew was back on the ship. Apollo 8 was over, a total success. This milestone in the space race had been won definitively by the U.S., and shortly thereafter the Soviets abandoned their Zond circumlunar project, judging it an anticlimax and admission of defeat to fly by the Moon after the Americans had already successfully orbited it.

This is the official NASA contemporary documentary about Apollo 8.

Here is an evening with the Apollo 8 astronauts recorded at the National Air and Space Museum on 2008-11-13 to commemorate the fortieth anniversary of the flight.

This is a reunion of the Apollo 8 astronauts on 2009-04-23.

As of this writing, all of the crew of Apollo 8 are alive, and, in a business where divorce was common, remain married to the women they wed as young military officers.

December 2018 Permalink

Kondo, Yoji, Frederick Bruhweiler, John Moore, and Charles Sheffield eds. Interstellar Travel and Multi-Generation Space Ships. Burlington, Ontario, Canada: Apogee Books, 2003. ISBN 1-896522-99-8.
This book is a collection of papers presented at a symposium organised in 2002 by the American Association for the Advancement of Science. More than half of the content discusses the motivations, technology, and prospects for interstellar flight (both robotic probes and “generation ship” exploration and colonisation missions), while the balance deals with anthropological, genetic, and linguistic issues in crew composition for a notional mission with a crew of 200 with a flight time of two centuries. An essay by Freeman Dyson on “Looking for Life in Unlikely Places” explores the signatures of ubiquitous vacuum-adapted life and how surprisingly easy it might be to detect, even as far as one light-year from Earth.

This volume contains the last published works of Charles Sheffield and Robert L. Forward, both of whom died in 2002. The papers are all accessible to the scientifically literate layman and, with one exception, of high quality. Regrettably, nobody seemed to have informed the linguist contributor that any interstellar mission would certainly receive a steady stream of broadband transmissions from the home planet (who would fund a multi-terabuck mission without the ability to monitor it and receive the results?), but that chapter is only four pages and may be deemed comic relief.

June 2007 Permalink

Kraft, Christopher C. Flight: My Life in Mission Control. New York: Dutton, 2001. ISBN 0-525-94571-7.

May 2001 Permalink

Kranz, Gene. Failure Is Not an Option. New York: Simon & Schuster, 2000. ISBN 0-7432-0079-9.

April 2001 Permalink

Launius, Roger D. and Dennis R. Jenkins. Coming Home. Washington: National Aeronautics and Space Administration, 2012. ISBN 978-0-16-091064-7. NASA SP-2011-593.
In the early decades of the twentieth century, when visionaries such as Konstantin Tsiolkovsky, Hermann Oberth, and Robert H. Goddard started to think seriously about how space travel might be accomplished, most of the focus was on how rockets might be designed and built which would enable their payloads to be accelerated to reach the extreme altitude and velocity required for long-distance ballistic or orbital flight. This is a daunting problem. The Earth has a deep gravity well: so deep that to place a satellite in a low orbit around it, you must not only lift the satellite from the Earth's surface to the desired orbital altitude (which isn't particularly difficult), but also impart sufficient velocity to it so that it does not fall back but, instead, orbits the planet. It's the speed that makes it so difficult.

Recall that the kinetic energy of a body is given by ½mv². If mass (m) is given in kilograms and velocity (v) in metres per second, energy is measured in joules. Note that the square of the velocity appears in the formula: if you triple the velocity, you need nine times the energy to accelerate the mass to that speed. A satellite must have a velocity of around 7.8 kilometres/second to remain in a low Earth orbit. This is about eight times the muzzle velocity of the 5.56×45mm NATO round fired by the M-16 and AR-15 rifles. Consequently, the satellite has sixty-four times the energy per unit mass of the rifle bullet, and the rocket which places it into orbit must expend all of that energy to launch it.

Every kilogram of a satellite in a low orbit has a kinetic energy of around 30 megajoules (thirty million joules). By comparison, the energy released by detonating a kilogram of TNT is 4.7 megajoules. The satellite, purely due to its motion, has more than six times the energy as an equal mass of TNT. The U.S. Space Shuttle orbiter had a mass, without payload, of around 70,000 kilograms. When preparing to leave orbit and return to Earth, its kinetic energy was about that of half a kiloton of TNT. During the process of atmospheric reentry and landing, in about half an hour, all of that energy must be dissipated in a non-destructive manner, until the orbiter comes to a stop on the runway with kinetic energy zero.

This is an extraordinarily difficult problem, which engineers had to confront as soon as they contemplated returning payloads from space to the Earth. The first payloads were, of course, warheads on intercontinental ballistic missiles. While these missiles did not go into orbit, they achieved speeds which were sufficiently fast as to present essentially the same problems as orbital reentry. When the first reconnaissance satellites were developed by the U.S. and the Soviet Union, the technology to capture images electronically and radio them to ground stations did not yet exist. The only option was to expose photographic film in orbit then physically return it to Earth for processing and interpretation. This was the requirement which drove the development of orbital reentry. The first manned orbital capsules employed technology proven by film return spy satellites. (In the case of the Soviets, the basic structure of the Zenit reconnaissance satellites and manned Vostok capsules was essentially the same.)

This book chronicles the history and engineering details of U.S. reentry and landing technology, for both unmanned and manned spacecraft. While many in the 1950s envisioned sleek spaceplanes as the vehicle of choice, when the time came to actually solve the problems of reentry, a seemingly counterintuitive solution came to the fore: the blunt body. We're all acquainted with the phenomenon of air friction: the faster an airplane flies, the hotter its skin gets. The SR-71, which flew at three times the speed of sound, had to be made of titanium since aluminium would have lost its strength at the temperatures which resulted from friction. But at the velocity of a returning satellite, around eight times faster than an SR-71, air behaves very differently. The satellite is moving so fast that air can't get out of the way and piles up in front of it. As the air is compressed, its temperature rises until it equals or exceeds that of the surface of the Sun. This heat is then radiated in all directions. That impinging upon the reentering body can, if not dealt with, destroy it.

A streamlined shape will cause the compression to be concentrated at the nose, leading to extreme heating. A blunt body, however, will cause a shock wave to form which stands off from its surface. Since the compressed air radiates heat in all directions, only that radiated in the direction of the body will be absorbed; the rest will be harmlessly radiated away into space, reducing total heating. There is still, however, plenty of heat to worry about.

Let's consider the Mercury capsules in which the first U.S. astronauts flew. They reentered blunt end first, with a heat shield facing the air flow. Compression in the shock layer ahead of the heat shield raised the air temperature to around 5800° K, almost precisely the surface temperature of the Sun. Over the reentry, the heat pulse would deposit a total of 100 megajoules per square metre of heat shield. The astronaut was just a few centimetres from the shield, and the temperature on the back side of the shield could not be allowed to exceed 65° C. How in the world do you accomplish that?

Engineers have investigated a wide variety of ways to beat the heat. The simplest are completely passive systems: they have no moving parts. An example of a passive system is a “heat sink”. You simply have a mass of some substance with high heat capacity (which means it can absorb a large amount of energy with a small rise in temperature), usually a metal, which absorbs the heat during the pulse, then slowly releases it. The heat sink must be made of a material which doesn't melt or corrode during the heat pulse. The original design of the Mercury spacecraft specified a beryllium heat sink design, and this was flown on the two suborbital flights, but was replaced for the orbital missions. The Space Shuttle used a passive heat shield of a different kind: ceramic tiles which could withstand the heat on their surface and provided insulation which prevented the heat from reaching the aluminium structure beneath. The tiles proved very difficult to manufacture, were fragile, and required a great deal of maintenance, but they were, in principle, reusable.

The most commonly used technology for reentry is ablation. A heat shield is fabricated of a material which, when subjected to reentry heat, chars and releases gases. The gases carry away the heat, while the charred material which remains provides insulation. A variety of materials have been used for ablative heat shields, from advanced silicone and carbon composites to oak wood, on some early Soviet and Chinese reentry experiments. Ablative heat shields were used on Mercury orbital capsules, in projects Gemini and Apollo, all Soviet and Chinese manned spacecraft, and will be used by the SpaceX and Boeing crew transport capsules now under development.

If the heat shield works and you make it through the heat pulse, you're still falling like a rock. The solution of choice for landing spacecraft has been parachutes, and even though they seem simple conceptually, in practice there are many details which must be dealt with, such as stabilising the falling craft so it won't tumble and tangle the parachute suspension lines when the parachute is deployed, and opening the canopy in multiple stages to prevent a jarring shock which might damage the parachute or craft.

The early astronauts were pilots, and never much liked the idea of having to be fished out of the ocean by the Navy at the conclusion of their flights. A variety of schemes were explored to allow piloted flight to a runway landing, including inflatable wings and paragliders, but difficulties developing the technologies and schedule pressure during the space race caused the Gemini and Apollo projects to abandon them in favour of parachutes and a splashdown. Not until the Space Shuttle were precision runway landings achieved, and now NASA has abandoned that capability. SpaceX hopes to eventually return their Crew Dragon capsule to a landing pad with a propulsive landing, but that is not discussed here.

In the 1990s, NASA pursued a variety of spaceplane concepts: the X-33, X-34, and X-38. These projects pioneered new concepts in thermal protection for reentry which would be less expensive and maintenance-intensive than the Space Shuttle's tiles. In keeping with NASA's practice of the era, each project was cancelled after consuming a large sum of money and extensive engineering development. The X-37 was developed by NASA, and when abandoned, was taken over by the Air Force, which operates it on secret missions. Each of these projects is discussed here.

This book is the definitive history of U.S. spacecraft reentry systems. There is a wealth of technical detail, and some readers may find there's more here than they wanted to know. No specialised knowledge is required to understand the descriptions: just patience. In keeping with NASA tradition, quaint units like inches, pounds, miles per hour, and British Thermal Units are used in most of the text, but then in the final chapters, the authors switch back and forth between metric and U.S. customary units seemingly at random. There are some delightful anecdotes, such as when the designers of NASA's new Orion capsule had to visit the Smithsonian's National Air and Space Museum to examine an Apollo heat shield to figure out how it was made, attached to the spacecraft, and the properties of the proprietary ablative material it employed.

As a NASA publication, this book is in the public domain. The paperback linked to above is a republication of the original NASA edition. The book may be downloaded for free from the book's Web page in three electronic formats: PDF, MOBI (Kindle), and EPUB. Get the PDF! While the PDF is a faithful representation of the print edition, the MOBI edition is hideously ugly and mis-formatted. Footnotes are interleaved in the text at random locations in red type (except when they aren't in red type), block quotes are not set off from the main text, dozens of hyphenated words and adjacent words are run together, and the index is completely useless: citing page numbers in the print edition which do not appear in the electronic edition; for some reason large sections of the index are in red type. I haven't looked at the EPUB edition, but given the lack of attention to detail evident in the MOBI, my expectations for it are not high.

April 2016 Permalink

Lawrie, Alan. Sacramento's Moon Rockets. Charleston, SC: Arcadia Publishing, 2015. ISBN 978-1-4671-3389-0.
In 1849 gold was discovered in California, setting off a gold rush which would bring a wave of prospectors and fortune seekers into one of the greatest booms in American history. By the early 20th century, the grizzled prospector panning for gold had given way to industrial extraction of the metal. In an age before anybody had heard the word “environmentalism”, this was accomplished in the most direct way possible: man made lakes were created on gold-bearing land, then a barge would dredge up the bottom and mix it with mercury, which would form an amalgam with the gold. The gold could later be separated, purified, and sold.

The process effectively destroyed the land on which it was used. The topsoil was ripped out, vegetation killed, and the jumbled remains after extraction dumped in barren hills of tailings. Half a century later, the mined-out land was considered unusable for either agriculture or residential construction. Some described it as a “moonscape”.

It was perhaps appropriate that, in the 1960s, this stark terrain became home to the test stands on which the upper stage of NASA's Saturn rockets were developed and tested before flight. Every Saturn upper stage, including those which launched Apollo flights to the Moon, underwent a full-duration flight qualification firing there before being shipped to Florida for launch.

When the Saturn project was approved, Douglas Aircraft Company won the contract to develop the upper stage, which would be powered by liquid hydrogen and liquid oxygen (LH2/LOX) and have the ability to restart in space, allowing the Apollo spacecraft to leave Earth orbit on a trajectory bound for the Moon. The initial upper stage was called the S-IV, and was used as the second stage of the Saturn I launcher flown between 1961 and 1965 to demonstrate heavy lift booster operations and do development work related to the Apollo project. The S-IV used a cluster of six RL10 engines, at the time the largest operational LH2/LOX engine. The Saturn I had eight engines on its first stage and six engines on the S-IV. Given the reliability of rocket engines at the time, many engineers were dubious of getting fourteen engines to work on every launch (although the Saturn I did have a limited engine out capability). Skeptics called it “Cluster's last stand.”

The S-IV stages were manufactured at the Douglas plant in Huntington Beach, California, but there was no suitable location near the plant where they could be tested. The abandoned mining land near Sacramento had been acquired by Aerojet for rocket testing, and Douglas purchased a portion for its own use. The outsized S-IV stage was very difficult to transport by road, so the ability to ship it by water from southern California to the test site via San Francisco Bay and the Sacramento River was a major advantage of the location.

The operational launchers for Apollo missions would be the Saturn IB and Saturn V, with the Saturn IB used for Earth orbital missions and the Saturn V for Moon flights and launching space stations. An upgraded upper stage, the S-IVB, would be used by these launchers, as the second stage of the Saturn IB and the third stage of the Saturn V. (S-IVBs for the two launchers differed in details, but the basic configuration was the same.) The six RL-10 engines of the S-IV were replaced by a single much more powerful J-2 engine which had, by that time, become available.

The Sacramento test facility was modified to do development and preflight testing of the S-IVB, and proceeded to test every flight stage. No rocket firing is ever routine, and in 1965 and 1967 explosions destroyed an S-IV test article and a flight S-IVB stage which was scheduled to be used in Apollo 8. Fortunately, there were no casualties from these spectacular accidents, and they provided the first data on the effects of large scale LH2/LOX explosions which proved to be far more benign than had been feared. It had been predicted that a LH2/LOX explosion would produce a blast equal to 65% of the propellant mass of TNT when, in fact, the measured blast was just 5% TNT equivalent mass. It's nice to know, but an expensive way to learn.

This book is not a detailed history of the Sacramento test facility but rather a photo gallery showing the construction of the site; transportation of stages by sea, road, and later by the amazing Super Guppy airplane; testing of S-IV and S-IVB stages; explosions and their aftermath; and a visit to the site fifty years later. The photos have well-researched and informative captions.

When you think of the Apollo program, the Cape, Houston, Huntsville, and maybe Slidell come to mind, but rarely Sacramento. And yet every Apollo mission relied upon a rocket stage tested at the Rancho Cordova site near that city. Here is a part of the grandiose effort to go to the Moon you probably haven't seen before. The book is just 96 pages and expensive (a small print run and colour on almost every page will do that), but there are many pictures collected here I've seen nowhere else.

September 2015 Permalink

Leinbach, Michael D. and Jonathan H. Ward. Bringing Columbia Home. New York: Arcade Publishing, [2018] 2020. ISBN 978-1-948924-61-0.
Author Michael Leinbach was Launch Director at the Kennedy Space Center when space shuttle orbiter Columbia was lost during its return to Earth on February 1st, 2003. In this personal account, he tells the story of locating, recovering, and reconstructing the debris from the orbiter, searching for and finding the remains of the crew, and learning the lessons, technical and managerial, from the accident.

April 2020 Permalink

Light, Michael and Andrew Chaikin. Full Moon. New York: Alfred A. Knopf, 1999. ISBN 0-375-40634-4.

July 2002 Permalink

Linenger, Jerry M. Off the Planet. New York: McGraw-Hill, 2000. ISBN 0-07-137230-X.

November 2001 Permalink

Mallan, Lloyd. Russia and the Big Red Lie. Greenwich, CT: Fawcett, 1959. LCCN 59004006.
It is difficult for those who did not live through the era to appreciate the extent to which Sputnik shook the self-confidence of the West and defenders of the open society and free markets around the world. If the West's social and economic systems were genuinely superior to totalitarian rule and central planning, then how had the latter, starting from a base only a half century before where illiterate peasants were bound to the land as serfs, and in little more than a decade after their country was devastated in World War II, managed to pull off a technological achievement which had so far eluded the West and was evidence of a mastery of rocketry which could put the United States heartland at risk? Suddenly the fellow travellers and useful idiots in the West were energised: “Now witness the power of this fully armed and operational socialist economy!”

The author, a prolific writer on aerospace and technology, was as impressed as anybody else by the stunning Soviet accomplishment, and undertook the daunting task of arranging a visit to the Soviet Union to see for himself the prowess of Soviet science and technology. After a halting start, he secured a visa and introductions from prominent U.S. scientists to their Soviet counterparts, and journeyed to the Soviet Union in April of 1958, travelled extensively in the country, visiting, among other destinations, Moscow, Leningrad, Odessa, Yalta, Krasnodar, Rostov-on-Don, Yerevan, Kharkov, and Alma-Ata, leaving Soviet soil in June 1958. He had extensive, on the record, meetings with a long list of eminent Soviet scientists and engineers, many members of the Soviet Academy of Sciences. And he came back with a conclusion utterly opposed to that of the consensus in the West: Soviet technological prowess was about 1% military-style brute force and 99% bluff and hoax.

As one intimately acquainted with Western technology, what he saw in the Soviet Union was mostly comparable to the state of the art in the West a decade earlier, and in many cases obviously copied from Western equipment. The scientists he interviewed, who had been quoted in the Soviet press as forecasting stunning achievements in the near future, often, when interviewed in person, said “that's all just theory—nobody is actually working on that”. The much-vaunted Soviet jet and turboprop airliners he'd heard of were nowhere in evidence anywhere he travelled, and evidence suggested that Soviet commercial aviation lacked navigation and instrument landing systems which were commonplace in the West.

Faced with evidence that Soviet technological accomplishments were simply another front in a propaganda offensive aimed at persuading the world of the superiority of communism, the author dug deeper into the specifics of Soviet claims, and here (from the perspective of half a century on) he got some things right and goofed on others. He goes to great length to argue that the Luna 1 Moon probe was a total hoax, based both on Soviet technological capability and the evidence of repeated failure by Western listening posts to detect its radio signals. Current thinking is that Luna 1 was a genuine mission intended to impact on the Moon, but the Soviet claim it was deliberately launched into solar orbit as an “artificial planet” propaganda aimed at covering up its missing the Moon due to a guidance failure. (This became obvious to all when the near-identical Luna 2 impacted the moon eight months later.) The fact that the Soviets possessed the technology to conduct lunar missions was demonstrated when Luna 3 flew around the Moon in October 1959 and returned the first crude images of its far side (other Luna 3 images). Although Mallan later claimed these images were faked and contained brush strokes, we now know they were genuine, since they are strikingly similar to subsequent imagery, including the albedo map from the Clementine lunar orbiter. “Vas you dere, Ivan?” Well, actually, yes. Luna 3 was the “boomerang” mission around the Moon which Mallan had heard of before visiting the Soviet Union but was told was just a theory when he was there. And yet, had the Soviets had the ability to communicate with Luna 1 at the distance of the Moon, there would have been no reason to make Luna 3 loop around the Moon in order to transmit its pictures from closer to the Earth—enigmas, enigmas, enigmas.

In other matters, the author is dead on, where distinguished Western “experts” and “analysts” were completely taken in by the propaganda. He correctly identifies the Soviet “ICBM” from the 1957 Red Square parade as an intermediate range missile closer to the German V-2 than an intercontinental weapon. (The Soviet ICBM, the R-7, was indeed tested in 1957, but it was an entirely different design and could never have been paraded on a mobile launcher; it did not enter operational service until 1959.) He is also almost precisely on the money when he estimates the Soviet “ICBM arsenal” as on the order of half a dozen missiles, while the CIA was talking about hundreds of Soviet missiles aimed at the West and demagogues were ratcheting up rhetoric about a “missile gap”.

You don't read this for factual revelations: everything discussed here is now known much better, and there are many conclusions drawn in this text from murky contemporary evidence which have proven incorrect. But if you wish to immerse yourself in the Cold War and imagine yourself trying to figure it all out from the sketchy and distorted information coming from the adversary, it is very enlightening. One wishes more people had listened to Mallan—how much folly we might have avoided.

There is also wisdom in what he got wrong. Space spectaculars can be accomplished in a military manner by expending vast resources coercively taken from the productive sector on centrally-planned projects with narrow goals. Consequently, it isn't surprising a command economy such as that of the Soviet Union managed to achieve milestones in space (while failing to deliver adequate supplies of soap and toilet paper to workers toiling in their “paradise”). Indeed, in many ways, the U.S. Apollo program was even more centrally planned than its Soviet counterpart, and the pernicious example it set has damaged efforts to sustainably develop and exploit space ever since.

This “Fawcett Book” is basically an issue of Mechanix Illustrated containing a single long article. It even includes the usual delightful advertisements. This work is, of course, hopelessly out of print. Used copies are available, but often at absurdly elevated prices for what amounts to a pulp magazine. Is this work in the public domain and hence eligible to be posted on the Web? I don't know. It may well be: it was published before 1978, and unless its copyright was renewed in 1987 when its original 28 year term expired, it is public domain. Otherwise, as a publication by a “corporate author”, it will remain in copyright until 2079, which makes a mockery of the “limited Times to Authors” provision of the U.S. Constitution. If somebody can confirm this work is in the public domain, I'll scan it and make it available on the Web.

March 2012 Permalink

Mankins, John C. The Case for Space Solar Power. Houston: Virginia Edition, 2014. ISBN 978-0-9913370-0-2.
As world population continues to grow and people in the developing world improve their standard of living toward the level of residents of industrialised nations, demand for energy will increase enormously. Even taking into account anticipated progress in energy conservation and forecasts that world population will reach a mid-century peak and then stabilise, the demand for electricity alone is forecasted to quadruple in the century from 2000 to 2100. If electric vehicles shift a substantial part of the energy consumed for transportation from hydrocarbon fuels to electricity, the demand for electric power will be greater still.

Providing this electricity in an affordable, sustainable way is a tremendous challenge. Most electricity today is produced by burning fuels such as coal, natural gas, and petroleum; by nuclear fission reactors; and by hydroelectric power generated by dams. Quadrupling electric power generation by any of these means poses serious problems. Fossil fuels may be subject to depletion, pose environmental consequences both in extraction and release of combustion products into the atmosphere, and are distributed unevenly around the world, leading to geopolitical tensions between have and have-not countries. Uranium fission is a technology with few environmental drawbacks, but operating it in a safe manner is very demanding and requires continuous vigilance over the decades-long lifespan of a power station. Further, the risk exists that nuclear material can be diverted for weapons use, especially if nuclear power stations proliferate into areas which are politically unstable. Hydroelectric power is clean, generally reliable (except in the case of extreme droughts), and inexhaustible, but unfortunately most rivers which are suitable for its generation have already been dammed, and potential projects which might be developed are insufficient to meet the demand.

Well, what about those “sustainable energy” projects the environmentalists are always babbling about: solar panels, eagle shredders (wind turbines), and the like? They do generate energy without fuel, but they are not the solution to the problem. In order to understand why, we need to look into the nature of the market for electricity, which is segmented into two components, even though the current flows through the same wires. The first is “base load” power. The demand for electricity varies during the day, from day to day, and seasonally (for example, electricity for air conditioning peaks during the mid-day hours of summer). The base load is the electricity demand which is always present, regardless of these changes in demand. If you look at a long-term plot of electricity demand and draw a line through the troughs in the curve, everything below that line is base load power and everything above it is “peak” power. Base load power is typically provided by the sources discussed in the previous paragraph: hydrocarbon, nuclear, and hydroelectric. Because there is a continuous demand for the power they generate, these plants are designed to run non-stop (with excess capacity to cover stand-downs for maintenance), and may be complicated to start up or shut down. In Switzerland, for example, 56% of base load power is produced from hydroelectric plants and 39% from nuclear fission reactors.

The balance of electrical demand, peak power, is usually generated by smaller power plants which can be brought on-line and shut down quickly as demand varies. Peaking plants sell their power onto the grid at prices substantially higher than base load plants, which compensates for their less efficient operation and higher capital costs for intermittent operation. In Switzerland, most peak energy is generated by thermal plants which can burn either natural gas or oil.

Now the problem with “alternative energy” sources such as solar panels and windmills becomes apparent: they produce neither base load nor peak power. Solar panels produce electricity only during the day, and when the Sun is not obscured by clouds. Windmills, obviously, only generate when the wind is blowing. Since there is no way to efficiently store large quantities of energy (all existing storage technologies raise the cost of electricity to uneconomic levels), these technologies cannot be used for base load power, since they cannot be relied upon to continuously furnish power to the grid. Neither can they be used for peak power generation, since the times at which they are producing power may not coincide with times of peak demand. That isn't to say these energy sources cannot be useful. For example, solar panels on the roofs of buildings in the American southwest make a tremendous amount of sense since they tend to produce power at precisely the times the demand for air conditioning is greatest. This can smooth out, but not replace, the need for peak power generation on the grid.

If we wish to dramatically expand electricity generation without relying on fossil fuels for base load power, there are remarkably few potential technologies. Geothermal power is reliable and inexpensive, but is only available in a limited number of areas and cannot come close to meeting the demand. Nuclear fission, especially modern, modular designs is feasible, but faces formidable opposition from the fear-based community. If nuclear fusion ever becomes practical, we will have a limitless, mostly clean energy source, but after sixty years of research we are still decades away from an operational power plant, and it is entirely possible the entire effort may fail. The liquid fluoride thorium reactor, a technology demonstrated in the 1960s, could provide centuries of energy without the nuclear waste or weapons diversion risks of uranium-based nuclear power, but even if it were developed to industrial scale it's still a “nuclear reactor” and can be expected to stimulate the same hysteria as existing nuclear technology.

This book explores an entirely different alternative. Think about it: once you get above the Earth's atmosphere and sufficiently far from the Earth to avoid its shadow, the Sun provides a steady 1.368 kilowatts per square metre, and will continue to do so, non-stop, for billions of years into the future (actually, the Sun is gradually brightening, so on the scale of hundreds of millions of years this figure will increase). If this energy could be harvested and delivered efficiently to Earth, the electricity needs of a global technological civilisation could be met with a negligible impact on the Earth's environment. With present-day photovoltaic cells, we can convert 40% of incident sunlight to electricity, and wireless power transmission in the microwave band (to which the Earth's atmosphere is transparent, even in the presence of clouds and precipitation) has been demonstrated at 40% efficiency, with 60% end-to-end efficiency expected for future systems.

Thus, no scientific breakthrough of any kind is required to harvest abundant solar energy which presently streams past the Earth and deliver it to receiving stations on the ground which feed it into the power grid. Since the solar power satellites would generate energy 99.5% of the time (with short outages when passing through the Earth's shadow near the equinoxes, at which time another satellite at a different longitude could pick up the load), this would be base load power, with no fuel source required. It's “just a matter of engineering” to calculate what would be required to build the collector satellite, launch it into geostationary orbit (where it would stay above the same point on Earth), and build the receiver station on the ground to collect the energy beamed down by the satellite. Then, given a proposed design, one can calculate the capital cost to bring such a system into production, its operating cost, the price of power it would deliver to the grid, and the time to recover the investment in the system.

Solar power satellites are not a new idea. In 1968, Peter Glaser published a description of a system with photovoltaic electricity generation and microwave power transmission to an antenna on Earth; in 1973 he was granted U.S. patent 3,781,647 for the system. In the 1970s NASA and the Department of Energy conducted a detailed study of the concept, publishing a reference design in 1979 which envisioned a platform in geostationary orbit with solar arrays measuring 5 by 25 kilometres and requiring a monstrous space shuttle with payload of 250 metric tons and space factories to assemble the platforms. Design was entirely conventional, using much the same technologies as were later used in the International Space Station (ISS) (but for a structure twenty times its size). Given that the ISS has a cost estimated at US$ 150 billion, NASA's 1979 estimate that a complete, operational solar power satellite system comprising 60 power generation platforms and Earth-based infrastructure would cost (in 2014 dollars) between 2.9 and 8.7 trillion might be considered optimistic. Back then, a trillion dollars was a lot of money, and this study pretty much put an end to serious consideration of solar power satellites in the U.S.for almost two decades. In the late 1990s, NASA, realising that much progress has been made in many of the enabling technologies for space solar power, commissioned a “Fresh Look Study”, which concluded that the state of the art was still insufficiently advanced to make power satellites economically feasible.

In this book, the author, after a 25-year career at NASA, recounts the history of solar power satellites to date and presents a radically new design, SPS-ALPHA (Solar Power Satellite by means of Arbitrarily Large Phased Array), which he argues is congruent with 21st century manufacturing technology. There are two fundamental reasons previous cost estimates for solar power satellites have come up with such forbidding figures. First, space hardware is hideously expensive to develop and manufacture. Measured in US$ per kilogram, a laptop computer is around $200/kg, a Boeing 747 $1400/kg, and a smart phone $1800/kg. By comparison, the Space Shuttle Orbiter cost $86,000/kg and the International Space Station around $110,000/kg. Most of the exorbitant cost of space hardware has little to do with the space environment, but is due to its being essentially hand-built in small numbers, and thus never having the benefit of moving down the learning curve as a product is put into mass production nor of automation in manufacturing (which isn't cost-effective when you're only making a few of a product). Second, once you've paid that enormous cost per kilogram for the space hardware, you have launch it from the Earth into space and transport it to the orbit in which it will operate. For communication satellites which, like solar power satellites, operate in geostationary orbit, current launchers cost around US$ 50,000 per kilogram delivered there. New entrants into the market may substantially reduce this cost, but without a breakthrough such as full reusability of the launcher, it will stay at an elevated level.

SPS-ALPHA tackles the high cost of space hardware by adopting a “hyper modular” design, in which the power satellite is composed of huge numbers of identical modules of just eight different types. Each of these modules is on a scale which permits prototypes to be fabricated in facilities no more sophisticated than university laboratories and light enough they fall into the “smallsat” category, permitting inexpensive tests in the space environment as required. A production power satellite, designed to deliver 2 gigawatts of electricity to Earth, will have almost four hundred thousand of each of three types of these modules, assembled in space by 4,888 robot arm modules, using more than two million interconnect modules. These are numbers where mass production economies kick in: once the module design has been tested and certified you can put it out for bids for serial production. And a factory which invests in making these modules inexpensively can be assured of follow-on business if the initial power satellite is a success, since there will a demand for dozens or hundreds more once its practicality is demonstrated. None of these modules is remotely as complicated as an iPhone, and once they are made in comparable quantities shouldn't cost any more. What would an iPhone cost if they only made five of them?

Modularity also requires the design to be distributed and redundant. There is no single-point failure mode in the system. The propulsion and attitude control module is replicated 200 times in the full design. As modules fail, for whatever cause, they will have minimal impact on the performance of the satellite and can be swapped out as part of routine maintenance. The author estimates than on an ongoing basis, around 3% of modules will be replaced per year.

The problem of launch cost is addressed indirectly by the modular design. Since no module masses more than 600 kg (the propulsion module) and none of the others exceed 100 kg, they do not require a heavy lift launcher. Modules can simply be apportioned out among a large number of flights of the most economical launchers available. Construction of a full scale solar power satellite will require between 500 and 1000 launches per year of a launcher with a capacity in the 10 to 20 metric ton range. This dwarfs the entire global launch industry, and will provide motivation to fund the development of new, reusable, launcher designs and the volume of business to push their cost down the learning curve, with a goal of reducing cost for launch to low Earth orbit to US$ 300–500 per kilogram. Note that the SpaceX Falcon Heavy, under development with a projected first flight in 2015, already is priced around US$ 1000/kg without reusability of the three core stages which is expected to be introduced in the future.

The author lays out five “Design Reference Missions” which progress from small-scale tests of a few modules in low Earth orbit to a full production power satellite delivering 2 gigawatts to the electrical grid. He estimates a cost of around US$ 5 billion to the pilot plant demonstrator and 20 billion to the first full scale power satellite. This is not a small sum of money, but is comparable to the approximately US$ 26 billion cost of the Three Gorges Dam in China. Once power satellites start to come on line, each feeding power into the grid with no cost for fuel and modest maintenance expenses (comparable to those for a hydroelectric dam), the initial investment does not take long to be recovered. Further, the power satellite effort will bootstrap the infrastructure for routine, inexpensive access to space, and the power satellite modules can also be used in other space applications (for example, very high power communication satellites).

The most frequently raised objection when power satellites are mentioned is fear that they could be used as a “death ray”. This is, quite simply, nonsense. The microwave power beam arriving at the Earth's surface will have an intensity between 10–20% of summer sunlight, so a mirror reflecting the Sun would be a more effective death ray. Extensive tests were done to determine if the beam would affect birds, insects, and aircraft flying through it and all concluded there was no risk. A power satellite which beamed down its power with a laser could be weaponised, but nobody is proposing that, since it would have problems with atmospheric conditions and cost more than microwave transmission.

This book provides a comprehensive examination of the history of the concept of solar power from space, the various designs proposed over the years and studies conducted of them, and an in-depth presentation of the technology and economic rationale for the SPS-ALPHA system. It presents an energy future which is very different from that which most people envision, provides a way to bring the benefits of electrification to developing regions without any environmental consequences whatever, and ensure a secure supply of electricity for the foreseeable future.

This is a rewarding, but rather tedious read. Perhaps it's due to the author's 25 years at NASA, but the text is cluttered with acronyms—there are fourteen pages of them defined in a glossary at the end of the book—and busy charts, some of which are difficult to read as reproduced in the Kindle edition. Copy editing is so-so: I noted 28 errors, and I wasn't especially looking for them. The index in the Kindle edition lists page numbers in the print edition which are useless because the electronic edition does not contain page numbers.

June 2014 Permalink

Manto, Cindy Donze. Michoud Assembly Facility. Charleston, SC: Arcadia Publishing, 2014. ISBN 978-1-5316-6969-0.
In March, 1763, King Louis XV of France made a land grant of 140 square kilometres to Gilbert Antoine St Maxent, the richest man in Louisiana Territory and commander of the militia. The grant required St Maxent to build a road across the swampy property, develop a plantation, and reserve all the trees in forested areas for the use of the French navy. When the Spanish took over the territory five years later, St Maxent changed his first names to “Gilberto Antonio” and retained title to the sprawling estate. In the decades that followed, the property changed hands and nations several times, eventually, now part of the United States, being purchased by another French immigrant, Antoine Michoud, who had left France after the fall of Napoleon, who his father had served as an official.

Michoud rapidly established himself as a prosperous businessman in bustling New Orleans, and after purchasing the large tract of land set about buying pieces which had been sold off by previous owners, re-assembling most of the original French land grant into one of the largest private land holdings in the United States. The property was mostly used as a sugar plantation, although territory and rights were ceded over the years for construction of a lighthouse, railroads, and telegraph and telephone lines. Much of the land remained undeveloped, and like other parts of southern Louisiana was a swamp or, as they now say, “wetlands”.

The land remained in the Michoud family until 1910, when it was sold in its entirety for US$410,000 in cash (around US$11 million today) to a developer who promptly defaulted, leading to another series of changes of ownership and dodgy plans for the land, which most people continued to refer to as the Michoud Tract. At the start of World War II, the U.S. government bought a large parcel, initially intended for construction of Liberty ships. Those plans quickly fell through, but eventually a huge plant was erected on the site which, starting in 1943, began to manufacture components for cargo aircraft, lifeboats, and components which were used in the Manhattan Project's isotope separation plants in Oak Ridge, Tennessee.

At the end of the war, the plant was declared surplus but, a few years later, with the outbreak of the Korean War, it was re-purposed to manufacture engines for Army tanks. It continued in that role until 1954 when it was placed on standby and, in 1958, once again declared surplus. There things stood until mid-1961 when NASA, charged by the new Kennedy administration to “put a man on the Moon” was faced with the need to build rockets in sizes and quantities never before imagined, and to do so on a tight schedule, racing against the Soviet Union.

In June, 1961, Wernher von Braun, director of the NASA Marshall Space Flight Center in Huntsville, Alabama, responsible for designing and building those giant boosters, visited the then-idle Michoud Ordnance Plant and declared it ideal for NASA's requirements. It had 43 acres (17 hectares) under one roof, the air conditioning required for precision work in the Louisiana climate, and was ready to occupy. Most critically, it was located adjacent to navigable waters which would allow the enormous rocket stages, far too big to be shipped by road, rail, or air, to be transported on barges to and from Huntsville for testing and Cape Canaveral in Florida to be launched.

In September 1961 NASA officially took over the facility, renaming it “Michoud Operations”, to be managed by NASA Marshall as the manufacturing site for the rockets they designed. Work quickly got underway to set up manufacturing of the first stage of the Saturn I and 1B rockets and prepare to build the much larger first stage of the Saturn V Moon rocket. Before long, new buildings dedicated to assembly and test of the new rockets, occupied both by NASA and its contractors, began to spring up around the original plant. In 1965, the installation was renamed the Michoud Assembly Facility, which name it bears to this day.

With the end of the Apollo program, it looked like Michoud might once again be headed for white elephant status, but the design selected for the Space Shuttle included a very large External Tank comparable in size to the first stage of the Saturn V which would be discarded on every flight. Michoud's fabrication and assembly facilities, and its access to shipping by barge were ideal for this component of the Shuttle, and a total of 135 tanks built at Michoud were launched on Shuttle missions between 1981 and 2011.

The retirement of the Space Shuttle once again put the future of Michoud in doubt. It was originally tapped to build the core stage of the Constellation program's Ares V booster, which was similar in size and construction to the Shuttle External Tank. The cancellation of Constellation in 2010 brought that to a halt, but then Congress and NASA rode to the rescue with the absurd-as-a-rocket but excellent-as-a-jobs-program Space Launch System (SLS), whose centre core stage also resembles the External Tank and Ares V. SLS first stage fabrication is presently underway at Michoud. Perhaps when the schedule-slipping, bugget-busting SLS is retired after a few flights (if, in fact, it ever flies at all), bringing to a close the era of giant taxpayer-funded throwaway rockets, the Michoud facility can be repurposed to more productive endeavours.

This book is largely a history of Michoud in photos and captions, with text introducing chapters on each phase of the facility's history. All of the photos are in black and white, and are well-reproduced. In the Kindle edition many can be expanded to show more detail. There are a number of copy-editing and factual errors in the text and captions, but not too many to distract or mislead the reader. The unidentified “visitors” shown touring the Michoud facility in July 1967 (chapter 3, Kindle location 392) are actually the Apollo 7 crew, Walter Schirra, Donn Eisele, and Walter Cunningham, who would fly on a Michoud-built Saturn 1B in October 1968.

For a book of just 130 pages, most of which are black and white photographs, the hardcover is hideously expensive (US$29 at this writing). The Kindle edition is still pricey (US$13 list price), but may be read for free by Kindle Unlimited subscribers.

June 2019 Permalink

McDonald, Allan J. and James R. Hansen. Truth, Lies, and O-Rings. Gainesville, FL: University Press of Florida, 2009. ISBN 978-0-8130-3326-6.
More than two decades have elapsed since Space Shuttle Challenger met its tragic end on that cold Florida morning in January 1986, and a shelf-full of books have been written about the accident and its aftermath, ranging from the five volume official report of the Presidential commission convened to investigate the disaster to conspiracy theories and accounts of religious experiences. Is it possible, at this remove, to say anything new about Challenger? The answer is unequivocally yes, as this book conclusively demonstrates.

The night before Challenger was launched on its last mission, Allan McDonald attended the final day before launch flight readiness review at the Kennedy Space Center, representing Morton Thiokol, manufacturer of the solid rocket motors, where he was Director of the Space Shuttle Solid Rocket Motor Project. McDonald initially presented Thiokol's judgement that the launch should be postponed because the temperatures forecast for launch day were far below the experience base of the shuttle program and an earlier flight at the lowest temperature to date had shown evidence of blow-by the O-ring seals in the solid rocket field joints. Thiokol engineers were concerned that low temperatures would reduce the resiliency of the elastomeric rings, causing them to fail to seal during the critical ignition transient. McDonald was astonished when NASA personnel, in a reversal of their usual rôle of challenging contractors to prove why their hardware was safe to fly, demanded that Thiokol prove the solid motor was unsafe in order to scrub the launch. Thiokol management requested a five minute offline caucus back at the plant in Utah (in which McDonald did not participate) which stretched to thirty minutes and ended up with a recommendation to launch. NASA took the unprecedented step of requiring a written approval to launch from Thiokol, which McDonald refused to provide, but which was supplied by his boss in Utah.

After the loss of the shuttle and its crew, and the discovery shortly thereafter that the proximate cause was almost certainly a leak in the aft field joint of the right solid rocket booster, NASA and Thiokol appeared to circle the wagons, trying to deflect responsibility from themselves and obscure the information available to decision makers in a position to stop the launch. It was not until McDonald's testimony to the Presidential Commission chaired by former Secretary of State William P. Rogers that the truth began to come out. This thrust McDonald, up to then an obscure engineering manager, into the media spotlight and the political arena, which he quickly discovered was not at all about his priorities as an engineer: finding out what went wrong and fixing it so it could never happen again.

This memoir, composed by McDonald from contemporary notes and documents with the aid of space historian James R. Hansen (author of the bestselling authorised biography of Neil Armstrong) takes the reader through the catastrophe and its aftermath, as seen by an insider who was there at the decision to launch, on a console in the firing room when disaster struck, before the closed and public sessions of the Presidential commission, pursued by sensation-hungry media, testifying before congressional committees, and consumed by the redesign and certification effort and the push to return the shuttle to flight. It is a personal story, but told in terms, as engineers are wont to do, based in the facts of the hardware, the experimental evidence, and the recollection of meetings which made the key decisions before and after the tragedy.

Anybody whose career may eventually land them, intentionally or not (the latter almost always the case), in the public arena can profit from reading this book. Even if you know nothing about and have no interest in solid rocket motors, O-rings, space exploration, or NASA, the dynamics of a sincere, dedicated engineer who was bent on doing the right thing encountering the ravenous media and preening politicians is a cautionary tale for anybody who finds themselves in a similar position. I wish I'd had the opportunity to read this book before my own Dark Night of the Soul encounter with a reporter from the legacy media. I do not mean to equate my own mild experience with the Hell that McDonald experienced—just to say that his narrative would have been a bracing preparation for what was to come.

The chapters on the Rogers Commission investigation provided, for me, a perspective I'd not previously encountered. Many people think of William P. Rogers primarily as Nixon's first Secretary of State who was upstaged and eventually replaced by Henry Kissinger. But before that Rogers was a federal prosecutor going after organised crime in New York City and then was Attorney General in the Eisenhower administration from 1957 to 1961. Rogers may have aged, but his skills as an interrogator and cross-examiner never weakened. In the sworn testimony quoted here, NASA managers, who come across like the kids who were the smartest in their high school class and then find themselves on the left side of the bell curve when they show up as freshmen at MIT, are pinned like specimen bugs to their own viewgraphs when they try to spin Rogers and his tag team of technical takedown artists including Richard Feynman, Neil Armstrong, and Sally Ride.

One thing which is never discussed here, but should be, is just how totally insane it is to use large solid rockets, in any form, in a human spaceflight program. Understand: solid rockets are best thought of as “directed bombs”, but if detonated at an inopportune time, or when not in launch configuration, can cause catastrophe. A simple spark of static electricity can suffice to ignite the propellant in a solid rocket, and once ignited there is no way to extinguish it until it is entirely consumed. Consider: in the Shuttle era, there are usually one or more Shuttle stacks in the Vehicle Assembly Building (VAB), and if NASA's Constellation Program continues, this building will continue to stack solid rocket motors in decades to come. Sooner or later, the inevitable is going to happen: a static spark, a crane dropping a segment, or an interference fit of two segments sending a hot fragment into the propellant below. The consequence: destruction of the VAB, all hardware inside, and the death of all people working therein. The expected stand-down of the U.S. human spaceflight program after such an event is on the order of a decade. Am I exaggerating the risks here? Well, maybe; you decide. But within two years, three separate disasters struck the production of large solid motors in 1985–1986. I shall predict: if NASA continue to use large solid motors in their human spaceflight program, there will be a decade-long gap in U.S. human spaceflight sometime in the next twenty years.

If you're sufficiently interested in these arcane matters to have read this far, you should read this book. Based upon notes, it's a bit repetitive, as many of the same matters were discussed in the various venues in which McDonald testified. But if you want to read a single book to prepare you for being unexpectedly thrust into the maw of ravenous media and politicians, I know of none better.

September 2009 Permalink

Miller, Roland. Abandoned in Place. Albuquerque: University of New Mexico Press, 2016. ISBN 978-0-8263-5625-3.
Between 1945 and 1970 humanity expanded from the surface of Earth into the surrounding void, culminating in 1969 with the first landing on the Moon. Centuries from now, when humans and their descendents populate the solar system and exploit resources dwarfing those of the thin skin and atmosphere of the home planet, these first steps may be remembered as the most significant event of our age, with all of the trivialities that occupy our quotidian attention forgotten. Not only were great achievements made, but grand structures built on Earth to support them; these may be looked upon in the future as we regard the pyramids or the great cathedrals.

Or maybe not. The launch pads, gantry towers, assembly buildings, test facilities, blockhouses, bunkers, and control centres were not built as monuments for the ages, but rather to accomplish time-sensitive goals under tight budgets, by the lowest bidder, and at the behest of a government famous for neglecting infrastructure. Once the job was done, the mission accomplished, the program concluded; the facilities that supported it were simply left at the mercy of the elements which, in locations like coastal Florida, immediately began to reclaim them. Indeed, half of the facilities pictured here no longer exist.

For more than two decades, author and photographer Roland Miller has been documenting this heritage before it succumbs to rust, crumbling concrete, and invasive vegetation. With unparalleled access to the sites, he has assembled this gallery of these artefacts of a great age of exploration. In a few decades, this may be all we'll have to remember them. Although there is rudimentary background information from a variety of authors, this is a book of photography, not a history of the facilities. In some cases, unless you know from other sources what you're looking at, you might interpret some of the images as abstract.

The hardcover edition is a “coffee table book”: large format and beautifully printed, with a corresponding price. The Kindle edition is, well, a Kindle book, and grossly overpriced for 193 pages with screen-resolution images and a useless index consisting solely of search terms.

A selection of images from the book may be viewed on the Abandoned in Place Web site.

May 2016 Permalink

Mullane, Mike. Riding Rockets. New York: Scribner, 2006. ISBN 0-7432-7682-5.
Mike Mullane joined NASA in 1978, one of the first group of astronauts recruited specifically for the space shuttle program. An Air Force veteran of 134 combat missions in Vietnam as back-seater in the RF-4C reconnaissance version of the Phantom fighter (imperfect eyesight disqualified him from pilot training), he joined NASA as a mission specialist and eventually flew on three shuttle missions: STS-41D in 1984, STS-27 in 1988, and STS-36 in 1990, the latter two classified Department of Defense missions for which he was twice awarded the National Intelligence Medal of Achievement. (Receipt of this medal was, at the time, itself a secret, but was declassified after the collapse of the Soviet Union. The work for which the medals were awarded remains secret to this day.)

As a mission specialist, Mullane never maneuvered the shuttle in space nor landed it on Earth, nor did he perform a spacewalk, mark any significant “first” in space exploration or establish any records apart from being part of the crew of STS-36 which flew the highest inclination (62°) orbit of any human spaceflight so far. What he has done here is write one of the most enlightening, enthralling, and brutally honest astronaut memoirs ever published, far and away the best describing the shuttle era. All of the realities of NASA in the 1980s which were airbrushed out by Public Affairs Officers with the complicity of an astronaut corps who knew that to speak to an outsider about what was really going on would mean they'd never get another flight assignment are dealt with head-on: the dysfunctional, intimidation- and uncertainty-based management culture, the gap between what astronauts knew about the danger and unreliability of the shuttle and what NASA was telling Congress and public, the conflict between battle-hardened military astronauts and perpetual student post-docs recruited as scientist-astronauts, the shameless toadying to politicians, and the perennial over-promising of shuttle capabilities and consequent corner-cutting and workforce exhaustion. (Those of a libertarian bent might wish they could warp back in time, shake the author by the shoulders, and remind him, “Hey dude, you're working for a government agency!”)

The realities of flying a space shuttle mission are described without any of the sugar-coating or veiled references common in other astronaut accounts, and always with a sense of humour. The deep-seated dread of strapping into an experimental vehicle with four million pounds of explosive fuel and no crew escape system is discussed candidly, along with the fact that, while universally shared by astronauts, it was, of course, never hinted to outsiders, even passengers on the shuttle who were told it was a kind of very fast, high-flying airliner. Even if the shuttle doesn't kill you, there's still the toilet to deal with, and any curiosity you've had about that particular apparatus will not outlast your finishing this book (the on-orbit gross-out prank on p. 179 may be too much even for “South Park”). Barfing in space and the curious and little-discussed effects of microgravity on the male and female anatomy which may someday contribute mightily to the popularity of orbital tourism are discussed in graphic detail. A glossary of NASA jargon and acronyms is included but there is no index, which would be a valuable addition.

February 2006 Permalink

Murray, Charles and Catherine Bly Cox. Apollo. Burkittsville, MD: South Mountain Books, [1989, 2004] 2010. ISBN 978-0-9760008-0-8.
On November 5, 1958, NASA, only four months old at the time, created the Space Task Group (STG) to manage its manned spaceflight programs. Although there had been earlier military studies of manned space concepts and many saw eventual manned orbital flights growing out of the rocket plane projects conducted by NASA's predecessor, the National Advisory Committee for Aeronautics (NACA) and the U.S. Air Force, at the time of the STG's formation the U.S. had no formal manned space program. The initial group numbered 45 in all, including eight secretaries and “computers”—operators of electromechanical desk calculators, staffed largely with people from the NACA's Langley Research Center and initially headquartered there. There were no firm plans for manned spaceflight, no budget approved to pay for it, no spacecraft, no boosters, no launch facilities, no mission control centre, no astronauts, no plans to select and train them, and no experience either with human flight above the Earth's atmosphere or with more than a few seconds of weightlessness. And yet this team, the core of an effort which would grow to include around 400,000 people at NASA and its 20,000 industry and academic contractors, would, just ten years and nine months later, on July 20th, 1969, land two people on the surface of the Moon and then return them safely to the Earth.

Ten years is not a long time when it comes to accomplishing a complicated technological project. Development of the Boeing 787, a mid-sized commercial airliner which flew no further, faster, or higher than its predecessors, and was designed and built using computer-aided design and manufacturing technologies, took eight years from project launch to entry into service, and the F-35 fighter plane only entered service and then only in small numbers of one model a full twenty-three years after the start of its development.

In November, 1958, nobody in the Space Task Group was thinking about landing on the Moon. Certainly, trips to the Moon had been discussed in fables from antiquity to Jules Verne's classic De la terre à la lune of 1865, and in 1938 members of the British Interplanetary Society published a (totally impractical) design for a Moon rocket powered by more than two thousand solid rocket motors bundled together, which would be discarded once burned out, but only a year since the launch of the first Earth satellite and when nothing had been successfully returned from Earth orbit to the Earth, talk of manned Moon ships sounded like—lunacy.

The small band of stalwarts at the STG undertook the already daunting challenge of manned space flight with an incremental program they called Project Mercury, whose goal was to launch a single man into Earth orbit in a capsule (unable to change its orbit once released from the booster rocket, it barely deserved the term “spacecraft”) atop a converted Atlas intercontinental ballistic missile. In essence, the idea was to remove the warhead, replace it with a tiny cone-shaped can with a man in it, and shoot him into orbit. At the time the project began, the reliability of the Atlas rocket was around 75%, so NASA could expect around one in four launches to fail, with the Atlas known for spectacular explosions on the ground or on the way to space. When, in early 1960, the newly-chosen Mercury astronauts watched a test launch of the rocket they were to ride, it exploded less than a minute after launch. This was the fifth consecutive failure of an Atlas booster (although not all were so spectacular).

Doing things which were inherently risky on tight schedules with a shoestring budget (compared to military projects) and achieving an acceptable degree of safety by fanatic attention to detail and mountains of paperwork (NASA engineers quipped that no spacecraft could fly until the mass of paper documenting its construction and test equalled that of the flight hardware) became an integral part of the NASA culture. NASA was proceeding on its deliberate, step-by-step development of Project Mercury, and in 1961 was preparing for the first space flight by a U.S. astronaut, not into orbit on an Atlas, just a 15 minute suborbital hop on a version of the reliable Redstone rocket that launched the first U.S. satellite in 1958 when, on April 12, 1961, they were to be sorely disappointed when the Soviet Union launched Yuri Gagarin into orbit on Vostok 1. Not only was the first man in space a Soviet, they had accomplished an orbital mission, which NASA hadn't planned to attempt until at least the following year.

On May 5, 1961, NASA got back into the game, or at least the minor league, when Alan Shepard was launched on Mercury-Redstone 3. Sure, it was just a 15 minute up and down, but at least an American had been in space, if only briefly, and it was enough to persuade a recently-elected, young U.S. president smarting from being scooped by the Soviets to “take longer strides”. On May 25, less than three weeks after Shepard's flight, before a joint session of Congress, President Kennedy said, “I believe that this nation should commit itself to achieving the goal, before this decade is out, of landing a man on the Moon and returning him safely to Earth.” Kennedy had asked his vice president, Lyndon Johnson, what goal the U.S. could realistically hope to achieve before the Soviets, and after consulting with the NASA administrator, James Webb, a Texas oil man and lawyer, and no other NASA technical people other than Wernher von Braun, he reported that a manned Moon landing was the only milestone the Soviets, with their heavy boosters and lead in manned space flight, were unlikely to do first. So, to the Moon it was.

The Space Task Group people who were, ultimately going to be charged with accomplishing this goal and had no advance warning until they heard Kennedy's speech or got urgent telephone calls from colleagues who had also heard the broadcast were, in the words of their leader, Robert Gilruth, who had no more warning than his staff, “aghast”. He and his team had, like von Braun in the 1950s, envisioned a deliberate, step-by-step development of space flight capability: manned orbital flight, then a more capable spacecraft with a larger crew able to maneuver in space, a space station to explore the biomedical issues of long-term space flight and serve as a base to assemble craft bound farther into space, perhaps a reusable shuttle craft to ferry crew and cargo to space without (wastefully and at great cost) throwing away rockets designed as long-range military artillery on every mission,followed by careful reconnaissance of the Moon by both unmanned and manned craft to map its surface, find safe landing zones, and then demonstrate the technologies that would be required to get people there and back safely.

All that was now clearly out the window. If Congress came through with the massive funds it would require, going to the Moon would be a crash project like the Manhattan Project to build the atomic bomb in World War II, or the massive industrial mobilisation to build Liberty Ships or the B-17 and B-29 bombers. The clock was ticking: when Kennedy spoke, there were just 3142 days until December 31, 1969 (yes, I know the decade actually ends at the end of 1970, since there was no year 0 in the Gregorian calendar, but explaining this to clueless Americans is a lost cause), around eight years and seven months. What needed to be done? Everything. How much time was there to do it? Not remotely enough. Well, at least the economy was booming, politicians seemed willing to pay the huge bills for what needed to be done, and there were plenty of twenty-something newly-minted engineering graduates ready and willing to work around the clock without a break to make real what they'd dreamed of since reading science fiction in their youth.

The Apollo Project was simultaneously one of the most epochal and inspiring accomplishments of the human species, far more likely to be remembered a thousand years hence than anything else that happened in the twentieth century, and at the same time a politically-motivated blunder which retarded human expansion into the space frontier. Kennedy's speech was at the end of May 1961. Perhaps because the Space Task Group was so small, it and NASA were able to react with a speed which is stunning to those accustomed to twenty year development projects for hardware far less complicated than Apollo.

In June and July [1961], detailed specifications for the spacecraft hardware were completed. By the end of July, the Requests for Proposals were on the street.

In August, the first hardware contract was awarded to M.I.T.'s Instrumentation Laboratory for the Apollo guidance system. NASA selected Merritt Island, Florida, as the site for a new spaceport and acquired 125 square miles of land.

In September, NASA selected Michoud, Louisiana, as the production facility for the Saturn rockets, acquired a site for the Manned Spacecraft Center—the Space Task Group grown up—south of Houston, and awarded the contract for the second stage of the Saturn [V] to North American Aviation.

In October, NASA acquired 34 square miles for a Saturn test facility in Mississippi.

In November, the Saturn C-1 was successfully launched with a cluster of eight engines, developing 1.3 million pounds of thrust. The contract for the command and service module was awarded to North American Aviation.

In December, the contract for the first stage of the Saturn [V] was awarded to Boeing and the contract for the third stage was awarded to Douglas Aircraft.

By January of 1962, construction had begun at all of the acquired sites and development was under way at all of the contractors.

Such was the urgency with which NASA was responding to Kennedy's challenge and deadline that all of these decisions and work were done before deciding on how to get to the Moon—the so-called “mission mode”. There were three candidates: direct-ascent, Earth orbit rendezvous (EOR), and lunar orbit rendezvous (LOR). Direct ascent was the simplest, and much like idea of a Moon ship in golden age science fiction. One launch from Earth would send a ship to the Moon which would land there, then take off and return directly to Earth. There would be no need for rendezvous and docking in space (which had never been attempted, and nobody was sure was even possible), and no need for multiple launches per mission, which was seen as an advantage at a time when rockets were only marginally reliable and notorious for long delays from their scheduled launch time. The downside of direct-ascent was that it would require an enormous rocket: planners envisioned a monster called Nova which would have dwarfed the Saturn V eventually used for Apollo and required new manufacturing, test, and launch facilities to accommodate its size. Also, it is impossible to design a ship which is optimised both for landing under rocket power on the Moon and re-entering Earth's atmosphere at high speed. Still, direct-ascent seemed to involve the least number of technological unknowns. Ever wonder why the Apollo service module had that enormous Service Propulsion System engine? When it was specified, the mission mode had not been chosen, and it was made powerful enough to lift the entire command and service module off the lunar surface and return them to the Earth after a landing in direct-ascent mode.

Earth orbit rendezvous was similar to what Wernher von Braun envisioned in his 1950s popular writings about the conquest of space. Multiple launches would be used to assemble a Moon ship in low Earth orbit, and then, when it was complete, it would fly to the Moon, land, and then return to Earth. Such a plan would not necessarily even require a booster as large as the Saturn V. One might, for example, launch the lunar landing and return vehicle on one Saturn I, the stage which would propel it to the Moon on a second, and finally the crew on a third, who would board the ship only after it was assembled and ready to go. This was attractive in not requiring the development of a giant rocket, but required on-time launches of multiple rockets in quick succession, orbital rendezvous and docking (and in some schemes, refuelling), and still had the problem of designing a craft suitable both for landing on the Moon and returning to Earth.

Lunar orbit rendezvous was originally considered a distant third in the running. A single large rocket (but smaller than Nova) would launch two craft toward the Moon. One ship would be optimised for flight through the Earth's atmosphere and return to Earth, while the other would be designed solely for landing on the Moon. The Moon lander, operating only in vacuum and the Moon's weak gravity, need not be streamlined or structurally strong, and could be potentially much lighter than a ship able to both land on the Moon and return to Earth. Finally, once its mission was complete and the landing crew safely back in the Earth return ship, it could be discarded, meaning that all of the hardware needed solely for landing on the Moon need not be taken back to the Earth. This option was attractive, requiring only a single launch and no gargantuan rocket, and allowed optimising the lander for its mission (for example, providing better visibility to its pilots of the landing site), but it not only required rendezvous and docking, but doing it in lunar orbit which, if they failed, would strand the lander crew in orbit around the Moon with no hope of rescue.

After a high-stakes technical struggle, in the latter part of 1962, NASA selected lunar orbit rendezvous as the mission mode, with each landing mission to be launched on a single Saturn V booster, making the decision final with the selection of Grumman as contractor for the Lunar Module in November of that year. Had another mission mode been chosen, it is improbable in the extreme that the landing would have been accomplished in the 1960s.

The Apollo architecture was now in place. All that remained was building machines which had never been imagined before, learning to do things (on-time launches, rendezvous and docking in space, leaving spacecraft and working in the vacuum, precise navigation over distances no human had ever travelled before, and assessing all of the “unknown unknowns” [radiation risks, effects of long-term weightlessness, properties of the lunar surface, ability to land on lunar terrain, possible chemical or biological threats on the Moon, etc.]) and developing plans to cope with them.

This masterful book is the story of how what is possibly the largest collection of geeks and nerds ever assembled and directed at a single goal, funded with the abundant revenue from an economic boom, spurred by a geopolitical competition against the sworn enemy of liberty, took on these daunting challenges and, one by one, overcame them, found a way around, or simply accepted the risk because it was worth it. They learned how to tame giant rocket engines that randomly blew up by setting off bombs inside them. They abandoned the careful step-by-step development of complex rockets in favour of “all-up testing” (stack all of the untested pieces the first time, push the button, and see what happens) because “there wasn't enough time to do it any other way”. People were working 16–18–20 hours a day, seven days a week. Flight surgeons in Mission Control handed out “go and whoa pills”—amphetamines and barbiturates—to keep the kids on the console awake at work and asleep those few hours they were at home—hey, it was the Sixties!

This is not a tale of heroic astronauts and their exploits. The astronauts, as they have been the first to say, were literally at the “tip of the spear” and would not have been able to complete their missions without the work of almost half a million uncelebrated people who made them possible, not to mention the hundred million or so U.S. taxpayers who footed the bill.

This was not a straight march to victory. Three astronauts died in a launch pad fire the investigation of which revealed shockingly slapdash quality control in the assembly of their spacecraft and NASA's ignoring the lethal risk of fire in a pure oxygen atmosphere at sea level pressure. The second flight of the Saturn V was a near calamity due to multiple problems, some entirely avoidable (and yet the decision was made to man the next flight of the booster and send the crew to the Moon). Neil Armstrong narrowly escaped death in May 1968 when the Lunar Landing Research Vehicle he was flying ran out of fuel and crashed. And the division of responsibility between the crew in the spacecraft and mission controllers on the ground had to be worked out before it would be tested in flight where getting things right could mean the difference between life and death.

What can we learn from Apollo, fifty years on? Other than standing in awe at what was accomplished given the technology and state of the art of the time, and on a breathtakingly short schedule, little or nothing that is relevant to the development of space in the present and future. Apollo was the product of a set of circumstances which happened to come together at one point in history and are unlikely to ever recur. Although some of those who worked on making it a reality were dreamers and visionaries who saw it as the first step into expanding the human presence beyond the home planet, to those who voted to pay the forbidding bills (at its peak, NASA's budget, mostly devoted to Apollo, was more than 4% of all Federal spending; in recent years, it has settled at around one half of one percent: a national commitment to space eight times smaller as a fraction of total spending) Apollo was seen as a key battle in the Cold War. Allowing the Soviet Union to continue to achieve milestones in space while the U.S. played catch-up or forfeited the game would reinforce the Soviet message to the developing world that their economic and political system was the wave of the future, leaving decadent capitalism in the dust.

A young, ambitious, forward-looking president, smarting from being scooped once again by Yuri Gagarin's orbital flight and the humiliation of the débâcle at the Bay of Pigs in Cuba, seized on a bold stroke that would show the world the superiority of the U.S. by deploying its economic, industrial, and research resources toward a highly visible goal. And, after being assassinated two and a half years later, his successor, a space enthusiast who had directed a substantial part of NASA's spending to his home state and those of his political allies, presented the program as the legacy of the martyred president and vigorously defended it against those who tried to kill it or reduce its priority. The U.S. was in an economic boom which would last through most of the Apollo program until after the first Moon landing, and was the world's unchallenged economic powerhouse. And finally, the federal budget had not yet been devoured by uncontrollable “entitlement” spending and national debt was modest and manageable: if the national will was there, Apollo was affordable.

This confluence of circumstances was unique to its time and has not been repeated in the half century thereafter, nor is it likely to recur in the foreseeable future. Space enthusiasts who look at Apollo and what it accomplished in such a short time often err in assuming a similar program: government funded, on a massive scale with lavish budgets, focussed on a single goal, and based on special-purpose disposable hardware suited only for its specific mission, is the only way to open the space frontier. They are not only wrong in this assumption, but they are dreaming if they think there is the public support and political will to do anything like Apollo today. In fact, Apollo was not even particularly popular in the 1960s: only at one point in 1965 did public support for funding of human trips to the Moon poll higher than 50% and only around the time of the Apollo 11 landing did 50% of the U.S. population believe Apollo was worth what was being spent on it.

In fact, despite being motivated as a demonstration of the superiority of free people and free markets, Project Apollo was a quintessentially socialist space program. It was funded by money extracted by taxation, its priorities set by politicians, and its operations centrally planned and managed in a top-down fashion of which the Soviet functionaries at Gosplan could only dream. Its goals were set by politics, not economic benefits, science, or building a valuable infrastructure. This was not lost on the Soviets. Here is Soviet Minister of Defence Dmitriy Ustinov speaking at a Central Committee meeting in 1968, quoted by Boris Chertok in volume 4 of Rockets and People.

…the Americans have borrowed our basic method of operation—plan-based management and networked schedules. They have passed us in management and planning methods—they announce a launch preparation schedule in advance and strictly adhere to it. In essence, they have put into effect the principle of democratic centralism—free discussion followed by the strictest discipline during implementation.

This kind of socialist operation works fine in a wartime crash program driven by time pressure, where unlimited funds and manpower are available, and where there is plenty of capital which can be consumed or borrowed to pay for it. But it does not create sustainable enterprises. Once the goal is achieved, the war won (or lost), or it runs out of other people's money to spend, the whole thing grinds to a halt or stumbles along, continuing to consume resources while accomplishing little. This was the predictable trajectory of Apollo.

Apollo was one of the noblest achievements of the human species and we should celebrate it as a milestone in the human adventure, but trying to repeat it is pure poison to the human destiny in the solar system and beyond.

This book is a superb recounting of the Apollo experience, told mostly about the largely unknown people who confronted the daunting technical problems and, one by one, found solutions which, if not perfect, were good enough to land on the Moon in 1969. Later chapters describe key missions, again concentrating on the problem solving which went on behind the scenes to achieve their goals or, in the case of Apollo 13, get home alive. Looking back on something that happened fifty years ago, especially if you were born afterward, it may be difficult to appreciate just how daunting the idea of flying to the Moon was in May 1961. This book is the story of the people who faced that challenge, pulled it off, and are largely forgotten today.

Both the 1989 first edition and 2004 paperback revised edition are out of print and available only at absurd collectors' prices. The Kindle edition, which is based upon the 2004 edition with small revisions to adapt to digital reader devices is available at a reasonable price, as is an unabridged audio book, which is a reading of the 2004 edition. You'd think there would have been a paperback reprint of this valuable book in time for the fiftieth anniversary of the landing of Apollo 11 (and the thirtieth anniversary of its original publication), but there wasn't.

Project Apollo is such a huge, sprawling subject that no book can possibly cover every aspect of it. For those who wish to delve deeper, here is a reading list of excellent sources. I have read all of these books and recommend every one. For those I have reviewed, I link to my review; for others, I link to a source where you can obtain the book.

If you wish to commemorate the landing of Apollo 11 in a moving ceremony with friends, consider hosting an Evoloterra celebration.

July 2019 Permalink

O'Leary, Brian. The Making of an Ex-Astronaut. Boston: Houghton Mifflin, 1970. LCCN 70-112277.
This book is out of print. The link above will search for used copies at abebooks.com.

July 2003 Permalink

O'Neill, Gerard K. The High Frontier. Mojave, CA: Space Studies Institute, [1976, 1977, 1982, 1989] 2013. ISBN 978-0-688-03133-6.
In the tumultuous year of 1969, Prof. Gerard K. O'Neill of Princeton University was tapped to teach the large freshman physics course at that institution. To motivate talented students who might find the pace of the course tedious, he organised an informal seminar which would explore challenging topics to which the basic physics taught in the main course could be applied. For the first topic of the seminar he posed the question, “Is a planetary surface the right place for an expanding technological civilisation?”. So fascinating were the results of investigating this question that the seminar never made it to the next topic, and working out its ramifications would occupy the rest of O'Neill's life.

By 1974, O'Neill and his growing group of informal collaborators had come to believe not only that the answer to that 1969 question was a definitive “no”, but that a large-scale expansion of the human presence into space, using the abundant energy and material resources available outside the Earth's gravity well was not a goal for the distant future but rather something which could be accomplished using only technologies already proved or expected in the next few years (such as the NASA's space shuttle, then under development). Further, the budget to bootstrap the settlement of space until the point at which the space settlements were self-sustaining and able to expand without further support was on the order of magnitude of the Apollo project and, unlike Apollo, would have an economic pay-off which would grow exponentially as space settlements proliferated.

As O'Neill wrote, the world economy had just been hit by the first of what would be a series of “oil shocks”, which would lead to a massive transfer of wealth from productive, developed economies to desert despotisms whose significance to the world economy and geopolitics would be precisely zero did they not happen to sit atop a pool of fuel (which they lacked the ability to discover and produce). He soon realised that the key to economic feasibility of space settlements was using them to construct solar power satellites to beam energy back to Earth.

Solar power satellites are just barely economically viable if the material from which they are made must be launched from the Earth, and many design concepts assume a dramatic reduction in launch costs and super-lightweight structure and high efficiency solar cells for the satellites, which adds to their capital cost. O'Neill realised that the materials which make up around 99% of the mass of a solar power satellite are available on the Moon, and a space settlement, with access to lunar material at a small fraction of the cost of launching from Earth and the ability to fabricate the very large power satellite structures in weightlessness would reduce the cost of space solar power to well below electricity prices of the mid-1970s (which were much lower than those of today).

In this book, a complete architecture is laid out, starting with initial settlements of “only” 10,000 people in a sphere about half a kilometre in diameter, rotating to provide Earth-normal gravity at the equator. This would be nothing like what one thinks of as a “space station”: people would live in apartments at a density comparable to small towns on Earth, surrounded by vegetation and with a stream running around the equator of the sphere. Lunar material would provide radiation shielding and mirrors would provide sunlight and a normal cycle of day and night.

This would be just a first step, with subsequent settlements much larger and with amenities equal to or exceeding those of Earth. Once access to the resources of asteroids (initially those in near-Earth or Earth-crossing orbits, and eventually the main belt) was opened, the space economy's reliance on the Earth would be only for settlers and lightweight, labour-intensive goods which made more sense to import. (For example, it might be some time before a space settlement built its own semiconductor fabrication facility rather than importing chips from those on Earth.)

This is the future we could be living in today, but turned our backs upon. Having read this book shortly after it first came out, it is difficult to describe just how bracing this optimistic, expansive view of the future was in the 1970s, when everything was brown and the human prospect suddenly seemed constrained by limited resources, faltering prosperity, and shrinking personal liberty. The curious thing about re-reading it today is that almost nothing has changed. Forty years later, O'Neill's roadmap for the future is just as viable an option for a visionary society as it was when initially proposed, and technological progress and understanding of the space environment has only improved its plausibility. The International Space Station, although a multi-decade detour from true space settlements, provides a testbed where technologies for those settlements can be explored (for example, solar powered closed-cycle Brayton engines as an alternative to photovoltaics for power generation, and high-yield agricultural techniques in a closed-loop ecosystem).

The re-appearance of this book in an electronic edition is timely, as O'Neill's ideas and the optimism for a better future they inspired seem almost forgotten today. Many people assume there was some technological flaw in his argument or that an economic show-stopper was discovered, yet none was. It was more like the reaction O'Neill encountered when he first tried to get his ideas into print in 1972. One reviewer, recommending against publication, wrote, “No one else is thinking in these terms, therefore the ideas must be wrong.” Today, even space “visionaries” imagine establishing human settlements on the Moon, Mars, and among the asteroids, with space travel seen as a way to get to these destinations and sustain pioneer communities there. This is a vision akin to long sea voyages to settle distant lands. O'Neill's High Frontier is something very different and epochal: the expansion of a species which evolved on the surface of a planet into the space around it and eventually throughout the solar system, using the abundant solar energy and material resources available there. This is like life expanding from the sea where it originated onto the land. It is the next step in the human adventure, and it can begin, just as it could have in 1976, within a decade of a developed society committing to make it so.

For some reason the Kindle edition, at least when viewed with the iPad Kindle application, displays with tiny type. I found I had to increase the font size by four steps to render it easily readable. Since font size is a global setting, that means than if you view another book, it shows up with giant letters like a first grade reader. The illustrations are dark and difficult to interpret in the Kindle edition—I do not recall whether this was also the case in the paperback edition I read many years ago.

May 2013 Permalink

Outzen, James D., ed. The Dorian Files Revealed. Chantilly, VA: Center for the Study of National Reconnaissance, 2015. ISBN 978-1-937219-18-5.
We often think of the 1960s as a “can do” time, when technological progress, societal self-confidence, and burgeoning economic growth allowed attempting and achieving great things: from landing on the Moon, global communications by satellite, and mass continental and intercontinental transportation by air. But the 1960s were also a time, not just of conflict and the dissolution of the postwar consensus, but also of some grand-scale technological boondoggles and disasters. There was the XB-70 bomber and its companion F-108 fighter plane, the Boeing 2707 supersonic passenger airplane, the NERVA nuclear rocket, the TFX/F-111 swing-wing hangar queen aircraft, and plans for military manned space programs. Each consumed billions of taxpayer dollars with little or nothing to show for the expenditure of money and effort lavished upon them. The present volume, consisting of previously secret information declassified in July 2015, chronicles the history of the Manned Orbiting Laboratory, the U.S. Air Force's second attempt to launch its own astronauts into space to do military tasks there.

The creation of NASA in 1958 took the wind out of the sails of the U.S. military services, who had assumed it would be they who would lead on the road into space and in exploiting space-based assets in the interest of national security. The designation of NASA as a civilian aerospace agency did not preclude military efforts in space, and the Air Force continued with its X-20 Dyna-Soar, a spaceplane intended to be launched on a Titan rocket which would return to Earth and land on a conventional runway. Simultaneous with the cancellation of Dyna-Soar in December 1963, a new military space program, the Manned Orbiting Laboratory (MOL) was announced.

MOL would use a modified version of NASA's Gemini spacecraft to carry two military astronauts into orbit atop a laboratory facility which they could occupy for up to 60 days before returning to Earth in the Gemini capsule. The Gemini and laboratory would be launched by a Titan III booster, requiring only a single launch and no orbital rendezvous or docking to accomplish the mission. The purpose of the program was stated as to “evaluate the utility of manned space flight for military purposes”. This was a cover story or, if you like, a bald-faced lie.

In fact, MOL was a manned spy satellite, intended to produce reconnaissance imagery of targets in the Soviet Union, China, and the communist bloc in the visual, infrared, and radar bands, plus electronic information in much higher resolution than contemporary unmanned spy satellites. Spy satellites operating in the visual spectrum lost on the order of half their images to cloud cover. With a man on board, exposures would be taken only when skies were clear, and images could be compensated for motion of the spacecraft, largely eliminating motion blur. Further, the pilots could scan for “interesting” targets and photograph them as they appeared, and conduct wide-area ocean surveillance.

None of the contemporary drawings showed the internal structure of the MOL, and most people assumed it was a large pressurised structure for various experiments. In fact, most of it was an enormous telescope aimed at the ground, with a 72 inch (1.83 metre) mirror and secondary optics capable of very high resolution photography of targets on the ground. When this document was declassified in 2015, all references to its resolution capability were replaced with statements such as {better than 1 foot}. It is, in fact, a simple geometrical optics calculation to determine that the diffraction-limited resolution of a 1.83 metre mirror in the visual band is around 0.066 arc seconds. In a low orbit suited to imaging in detail, this would yield a resolution of around 4 cm (1.6 inches) as a theoretical maximum. Taking optical imperfections, atmospheric seeing, film resolution, and imperfect motion compensation into account, the actual delivered resolution would be about half this (8 cm, 3.2 inches). Once they state the aperture of the primary mirror, this is easy to work out, so they wasted a lot of black redaction ink in this document. And then, on page 102, they note (not redacted), “During times of crisis the MOL could be transferred from its nominal 80-mile orbit to one of approximately 200–300 miles. In this higher orbit the system would have access to all targets in the Soviet Bloc approximately once every three days and be able to take photographs at resolutions of about one foot.” All right, if they have one foot (12 inch) resolution at 200 miles, then they have 4.8 inch (12 cm) resolution at 80 miles (or, if we take 250 miles altitude, 3.8 inches [9.7 cm]), entirely consistent with my calculation from mirror aperture.

This document is a management, financial, and political history of the MOL program, with relatively little engineering detail. Many of the technological developments of the optical system were later used in unmanned reconnaissance satellite programs and remain secret. What comes across in the sorry history of this program, which, between December 1963 and its cancellation in June of 1969 burned through billions of taxpayer dollars, is that the budgeting, project management, and definition and pursuit of well-defined project goals was just as incompetent as the redaction of technical details discussed in the previous paragraph. There are almost Marx brothers episodes where Florida politicians attempted to keep jobs in their constituencies by blocking launches into polar orbit from Vandenberg Air Force Base while the Air Force could not disclose that polar orbits were essential to overflying targets in the Soviet Union because the reconnaissance mission of MOL was a black program.

Along with this history, a large collection of documents and pictures, all previously secret (and many soporifically boring) has been released. As a publication of the U.S. government, this work is in the public domain.

November 2015 Permalink

Page, Joseph T., II. Vandenberg Air Force Base. Charleston, SC: Arcadia Publishing, 2014. ISBN 978-1-4671-3209-1.
Prior to World War II, the sleepy rural part of the southern California coast between Santa Barbara and San Luis Obispo was best known as the location where, in September 1923, despite a lighthouse having been in operation at Arguello Point since 1901, the U.S. Navy suffered its worst peacetime disaster, when seven destroyers, travelling at 20 knots, ran aground at Honda Point, resulting in the loss of all seven ships and the deaths of 23 crewmembers. In the 1930s, following additional wrecks in the area, a lifeboat station was established in conjunction with the lighthouse.

During World War II, the Army acquired 92,000 acres (372 km²) in the area for a training base which was called Camp Cooke, after a cavalry general who served in the Civil War, in wars with Indian tribes, and in the Mexican-American War. The camp was used for training Army troops in a variety of weapons and in tank maneuvers. After the end of the war, the base was closed and placed on inactive status, but was re-opened after the outbreak of war in Korea to train tank crews. It was once again mothballed in 1953, and remained inactive until 1957, when 64,000 acres were transferred to the U.S. Air Force to establish a missile base on the West Coast, initially called Cooke Air Force Base, intended to train missile crews and also serve as the U.S.'s first operational intercontinental ballistic missile (ICBM) site. On October 4th, 1958, the base was renamed Vandenberg Air Force Base in honour of the late General Hoyt Vandenberg, former Air Force Chief of Staff and Director of Central Intelligence.

On December 15, 1958, a Thor intermediate range ballistic missile was launched from the new base, the first of hundreds of launches which would follow and continue up to the present day. Starting in September 1959, three Atlas ICBMs armed with nuclear warheads were deployed on open launch pads at Vandenberg, the first U.S. intercontinental ballistic missiles to go on alert. The Atlas missiles remained part of the U.S. nuclear force until their retirement in May 1964.

With the advent of Earth satellites, Vandenberg became a key part of the U.S. military and civil space infrastructure. Launches from Cape Canaveral in Florida are restricted to a corridor directed eastward over the Atlantic ocean. While this is fine for satellites bound for equatorial orbits, such as the geostationary orbits used by many communication satellites, a launch into polar orbit, preferred by military reconnaissance satellites and Earth resources satellites because it allows them to overfly and image locations anywhere on Earth, would result in the rockets used to launch them dropping spent stages on land, which would vex taxpayers to the north and hotheated Latin neighbours to the south.

Vandenberg Air Force Base, however, situated on a point extending from the California coast, had nothing to the south but open ocean all the way to Antarctica. Launching southward, satellites could be placed into polar or Sun synchronous orbits without disturbing anybody but the fishes. Vandenberg thus became the prime launch site for U.S. reconnaissance satellites which, in the early days when satellites were short-lived and returned film to the Earth, required a large number of launches. The Corona spy satellites alone accounted for 144 launches from Vandenberg between 1959 and 1972.

With plans in the 1970s to replace all U.S. expendable launchers with the Space Shuttle, facilities were built at Vandenberg (Space Launch Complex 6) to process and launch the Shuttle, using a very different architecture than was employed in Florida. The Shuttle stack would be assembled on the launch pad, protected by a movable building that would retract prior to launch. The launch control centre was located just 365 metres from the launch pad (as opposed to 4.8 km away at the Kennedy Space Center in Florida), so the plan in case of a catastrophic launch accident on the pad essentially seemed to be “hope that never happens”. In any case, after spending more than US$4 billion on the facilities, after the Challenger disaster in 1986, plans for Shuttle launches from Vandenberg were abandoned, and the facility was mothballed until being adapted, years later, to launch other rockets.

This book, part of the “Images of America” series, is a collection of photographs (all black and white) covering all aspects of the history of the site from before World War II to the present day. Introductory text for each chapter and detailed captions describe the items shown and their significance to the base's history. The production quality is excellent, and I noted only one factual error in the text (the names of crew of Gemini 5). For a book of just 128 pages, the paperback is very expensive (US$22 at this writing). The Kindle edition is still pricey (US$13 list price), but may be read for free by Kindle Unlimited subscribers.

December 2019 Permalink

Pendle, George. Strange Angel. New York: Harcourt, 2005. ISBN 978-0-15-603179-0.
For those who grew up after World War II “rocket science” meant something extremely difficult, on the very edge of the possible, pursued by the brightest of the bright, often at risk of death or dire injury. In the first half of the century, however, “rocket” was a pejorative, summoning images of pulp magazines full of “that Buck Rogers stuff”, fireworks that went fwoosh—flash—bang if all went well, and often in the other order when it didn't, with aspiring rocketeers borderline lunatics who dreamed of crazy things like travelling to the Moon but usually ended blowing things up, including, but not limited to, themselves.

This was the era in which John Whiteside “Jack” Parsons came of age. Parsons was born and spent most of his life in Pasadena, California, a community close enough to Los Angeles to participate in its frontier, “anything goes” culture, but also steeped in well-heeled old wealth, largely made in the East and seeking the perpetually clement climate of southern California. Parsons was attracted to things that went fwoosh and bang from the very start. While still a high school senior, he was hired by the Hercules Powder Company, and continued to support himself as an explosives chemist for the rest of his life. He never graduated from college, no less pursued an advanced degree, but his associates and mentors, including legends such as Theodore von Kármán were deeply impressed by his knowledge and meticulously careful work with dangerous substances and gave him their highest recommendations. On several occasions he was called as an expert witness to testify in high-profile trials involving bombings.

And yet, at the time, to speak seriously about rockets was as outré as to admit one was a fan of “scientifiction” (later science fiction), or a believer in magic. Parsons was all-in on all of them. An avid reader of science fiction and member of the Los Angeles Science Fantasy Society, Parsons rubbed shoulders with Ray Bradbury, Robert Heinlein, and Forrest J. Ackerman. On the darker side, Parsons became increasingly involved in the Ordo Templi Orientis (OTO), followers of Aleister Crowley, and practitioners of his “magick”. One gets the sense that Parsons saw no conflict whatsoever among these pursuits—all were ways to transcend the prosaic everyday life and explore a universe enormously larger and stranger than even that of Los Angeles and its suburbs.

Parsons and his small band of rocket enthusiasts, “the suicide squad”, formed an uneasy alliance with the aeronautical laboratory of the California Institute of Technology, and with access to their resources and cloak of respectability, pursued their dangerous experiments first on campus, and then after a few embarrassing misadventures, in Arroyo Seco behind Pasadena. With the entry of the United States into World War II, the armed services had difficult problems to solve which overcame the giggle factor of anything involving the word “rocket”. In particular, the U.S. Navy had an urgent need to launch heavily-laden strike aircraft from short aircraft carrier decks (steam catapults were far in the future), and were willing to consider even Buck Rogers rockets to get them off the deck. Well, at least as long as you didn't call them “rockets”! So, the Navy sought to procure “Jet Assisted Take-Off” units, and Caltech created the “Jet Propulsion Laboratory” with Parsons as a founder to develop them, and then its members founded the Aerojet Engineering Corporation to build them in quantity. Nope, no rockets around here, nowhere—just jets.

Even as Parsons' rocket dreams came true and began to make him wealthy, he never forsook his other interests: they were all integral to him. He advanced in Crowley's OTO, became a regular correspondent of the Great Beast, and proprietor of the OTO lodge in Pasadena, home to a motley crew of bohemians who prefigured the beatniks and hippies of the 1950s and '60s. And he never relinquished his interest in science fiction, taking author L. Ron Hubbard into his community. Hubbard, a world class grifter even in his early days, took off with Parsons' girlfriend and most of his savings on the promise of buying yachts in Florida and selling them at a profit in California. Uh-huh! I'd put it down to destructive engrams.

Amidst all of this turmoil, Parsons made one of the most important inventions in practical rocketry of the 20th century. Apart from the work of Robert Goddard, which occurred largely disconnected from others due to Goddard's obsessive secrecy due to his earlier humiliation by learned ignoramuses, and the work by the German rocket team, conducted in secrecy in Nazi Germany, rockets mostly meant solid rockets, and solid rockets were little changed from mediaeval China: tubes packed with this or that variant of black powder which went fwoosh all at once when ignited. Nobody before Parsons saw an alternative to this. When faced by the need for a reliable, storable, long-duration burn propellant for Navy JATO boosters, he came up with the idea of castable solid propellant (initially based upon asphalt and potassium perchlorate), which could be poured as a liquid into a booster casing with a grain shape which permitted tailoring the duration and thrust profile of the motor to the mission requirements. Every single solid rocket motor used today employs this technology, and Jack Parsons, high school graduate and self-taught propulsion chemist, invented it all by himself.

On June 17th, 1952, an explosion destroyed a structure on Pasadena's Orange Grove Avenue where Jack Parsons had set up his home laboratory prior to his planned departure with his wife to Mexico. He said he had just one more job to do for his client, a company producing explosives for Hollywood special effects. Parsons was gravely injured and pronounced dead at the hospital.

The life of Jack Parsons was one which could only have occurred in the time and place he lived it. It was a time when a small band of outcasts could have seriously imagined building a rocket and travelling to the Moon; a time when the community they lived in was aboil with new religions, esoteric cults, and alternative lifestyles; and an entirely new genre of fiction was exploring the ultimate limits of the destiny of humanity and its descendents. Jack swam in this sea and relished it. His short life (just 37 years) was lived in a time and place which has never existed before and likely will never exist again. The work he did, the people he influenced, and the consequences cast a long shadow still visible today (every time you see a solid rocket booster heave a launcher off the pad, its coruscant light, casting that shadow, is Jack Parsons' legacy). This is a magnificent account of a singular life which changed our world, and is commemorated on the rock next door. On the lunar far side the 40 kilometre diameter crater Parsons is named for the man who dreamt of setting foot, by rocketry or magick, upon that orb and, in his legacy, finally did with a big footprint indeed—more than eight times larger than the one named for that Armstrong fellow.

July 2012 Permalink

Pooley, Charles and Ed LeBouthillier. Microlaunchers. Seattle: CreateSpace, 2013. ISBN 978-1-4912-8111-6.
Many fields of engineering are subject to scaling laws: as you make something bigger or smaller various trade-offs occur, and the properties of materials, cost, or other design constraints set limits on the largest and smallest practical designs. Rockets for launching payloads into Earth orbit and beyond tend to scale well as you increase their size. Because of the cube-square law, the volume of propellant a tank holds increases as the cube of the size while the weight of the tank goes as the square (actually a bit faster since a larger tank will require more robust walls, but for a rough approximation calling it the square will do). Viable rockets can get very big indeed: the Sea Dragon, although never built, is considered a workable design. With a length of 150 metres and 23 metres in diameter, it would have more than ten times the first stage thrust of a Saturn V and place 550 metric tons into low Earth orbit.

What about the other end of the scale? How small could a space launcher be, what technologies might be used in it, and what would it cost? Would it be possible to scale a launcher down so that small groups of individuals, from hobbyists to college class projects, could launch their own spacecraft? These are the questions explored in this fascinating and technically thorough book. Little practical work has been done to explore these questions. The smallest launcher to place a satellite in orbit was the Japanese Lambda 4S with a mass of 9400 kg and length of 16.5 metres. The U.S. Vanguard rocket had a mass of 10,050 kg and length of 23 metres. These are, though small compared to the workhorse launchers of today, still big, heavy machines, far beyond the capabilities of small groups of people, and sufficiently dangerous if something goes wrong that they require launch sites in unpopulated regions.

The scale of launchers has traditionally been driven by the mass of the payload they carry to space. Early launchers carried satellites with crude 1950s electronics, while many of their successors were derived from ballistic missiles sized to deliver heavy nuclear warheads. But today, CubeSats have demonstrated that useful work can be done by spacecraft with a volume of one litre and mass of 1.33 kg or less, and the PhoneSat project holds out the hope of functional spacecraft comparable in weight to a mobile telephone. While to date these small satellites have flown as piggy-back payloads on other launches, the availability of dedicated launchers sized for them would increase the number of launch opportunities and provide access to trajectories unavailable in piggy-back opportunities.

Just because launchers have tended to grow over time doesn't mean that's the only way to go. In the 1950s and '60s many people expected computers to continue their trend of getting bigger and bigger to the point where there were a limited number of “computer utilities” with vast machines which customers accessed over the telecommunication network. But then came the minicomputer and microcomputer revolutions and today the computing power in personal computers and mobile devices dwarfs that of all supercomputers combined. What would it take technologically to spark a similar revolution in space launchers?

With the smallest successful launchers to date having a mass of around 10 tonnes, the authors choose two weight budgets: 1000 kg on the high end and 100 kg as the low. They divide these budgets into allocations for payload, tankage, engines, fuel, etc. based upon the experience of existing sounding rockets, then explore what technologies exist which might enable such a vehicle to achieve orbital or escape velocity. The 100 kg launcher is a huge technological leap from anything with which we have experience and probably could be built, if at all, only after having gained experience from earlier generations of light launchers. But then the current state of the art in microchip fabrication would have seemed like science fiction to researchers in the early days of integrated circuits and it took decades of experience and generation after generation of chips and many technological innovations to arrive where we are today. Consequently, most of the book focuses on a three stage launcher with the 1000 kg mass budget, capable of placing a payload of between 150 and 200 grams on an Earth escape trajectory.

The book does not spare the rigour. The reader is introduced to the rocket equation, formulæ for aerodynamic drag, the standard atmosphere, optimisation of mixture ratios, combustion chamber pressure and size, nozzle expansion ratios, and a multitude of other details which make the difference between success and failure. Scaling to the size envisioned here without expensive and exotic materials and technologies requires out of the box thinking, and there is plenty on display here, including using beverage cans for upper stage propellant tanks.

A 1000 kg space launcher appears to be entirely feasible. The question is whether it can be done without the budget of hundreds of millions of dollars and years of development it would certainly take were the problem assigned to an aerospace prime contractor. The authors hold out the hope that it can be done, and observe that hobbyists and small groups can begin working independently on components: engines, tank systems, guidance and navigation, and so on, and then share their work precisely as open source software developers do so successfully today.

This is a field where prizes may work very well to encourage development of the required technologies. A philanthropist might offer, say, a prize of a million dollars for launching a 150 gram communicating payload onto an Earth escape trajectory, and a series of smaller prizes for engines which met the requirements for the various stages, flight-weight tankage and stage structures, etc. That way teams with expertise in various areas could work toward the individual prizes without having to take on the all-up integration required for the complete vehicle.

This is a well-researched and hopeful look at a technological direction few have thought about. The book is well written and includes all of the equations and data an aspiring rocket engineer will need to get started. The text is marred by a number of typographical errors (I counted two dozen) but only one trivial factual error. Although other references are mentioned in the text, a bibliography of works for those interested in exploring further would be a valuable addition. There is no index.

January 2014 Permalink

Portree, David S. F. Humans to Mars. Washington: National Aeronautics and Space Administration, 2001. NASA SP-2001-4521.
Ever since, in the years following World War II, people began to think seriously about the prospects for space travel, visionaries have looked beyond the near-term prospects for flights into Earth orbit, space stations, and even journeys to the Moon, toward the red planet: Mars. Unlike Venus, eternally shrouded by clouds, or the other planets which were too hot or cold to sustain life as we know it, Mars, about half the size of the Earth, had an atmosphere, a day just a little longer than the Earth's, seasons, and polar caps which grew and shrank with the seasons. There were no oceans, but water from the polar caps might sustain life on the surface, and there were dark markings which appeared to change during the Martian year, which some interpreted as plant life that flourished as polar caps melted in the spring and receded as they grew in the fall.

In an age where we have high-resolution imagery of the entire planet, obtained from orbiting spacecraft, telescopes orbiting Earth, and ground-based telescopes with advanced electronic instrumentation, it is often difficult to remember just how little was known about Mars in the 1950s, when people first started to think about how we might go there. Mars is the next planet outward from the Sun, so its distance and apparent size vary substantially depending upon its relative position to Earth in their respective orbits. About every two years, Earth “laps” Mars and it is closest (“at opposition”) and most easily observed. But because the orbit of Mars is elliptic, its distance varies from one opposition to the next, and it is only every 15 to 17 years that a near-simultaneous opposition and perihelion render Mars most accessible to Earth-based observation.

But even at a close opposition, Mars is a challenging telescopic target. At a close encounter, such as the one which will occur in the summer of 2018, Mars has an apparent diameter of only around 25 arc seconds. By comparison, the full Moon is about half a degree, or 1800 arc seconds: 72 times larger than Mars. To visual observers, even at a favourable opposition, Mars is a difficult object. Before the advent of electronic sensors in the 1980s, it was even more trying to photograph. Existing photographic film and plates were sufficiently insensitive that long exposures, measured in seconds, were required, and even from the best observing sites, the turbulence in the Earth's atmosphere smeared out details, leaving only the largest features recognisable. Visual observers were able to glimpse more detail in transient moments of still air, but had to rely upon their memory to sketch them. And the human eye is subject to optical illusions, seeing patterns where none exist. Were the extended linear features called “canals” real? Some observers saw and sketched them in great detail, while others saw nothing. Photography could not resolve the question.

Further, the physical properties of the planet were largely unknown. If you're contemplating a mission to land on Mars, it's essential to know the composition and density of its atmosphere, the temperatures expected at potential landing sites, and the terrain which a lander would encounter. None of these were known much beyond the level of educated guesses, which turned out to be grossly wrong once spacecraft probe data started to come in.

But ignorance of the destination didn't stop people from planning, or at least dreaming. In 1947–48, Wernher von Braun, then working with the U.S. Army at the White Sands Missile Range in New Mexico, wrote a novel called The Mars Project based upon a hypothetical Mars mission. A technical appendix presented detailed designs of the spacecraft and mission. While von Braun's talent as an engineer was legendary, his prowess as a novelist was less formidable, and the book never saw print, but in 1952 the appendix was published by itself.

One thing of which von Braun was never accused was thinking small, and in this first serious attempt to plan a Mars mission, he envisioned something more like an armada than the lightweight spacecraft we design today. At a time when the largest operational rocket, the V-2, had a payload of just one tonne, which it could throw no further than 320 km on a suborbital trajectory, von Braun's Mars fleet would consist of ten ships, each with a mass of 4,000 tons, and a total crew of seventy. The Mars ships would be assembled in orbit from parts launched on 950 flights of reusable three-stage ferry rockets. To launch all of the components of the Mars fleet and the fuel they would require would burn a total of 5.32 million tons of propellant in the ferry ships. Note that when von Braun proposed this, nobody had ever flown even a two stage rocket, and it would be ten years before the first unmanned Earth satellite was launched.

Von Braun later fleshed out his mission plans for an illustrated article in Collier's magazine as part of their series on the future of space flight. Now he envisioned assembling the Mars ships at the toroidal space station in Earth orbit which had figured in earlier installments of the series. In 1956, he published a book co-authored with Willy Ley, The Exploration of Mars, in which he envisioned a lean and mean expedition with just two ships and a crew of twelve, which would require “only” four hundred launches from Earth to assemble, provision, and fuel.

Not only was little understood about the properties of the destination, nothing at all was known about what human crews would experience in space, either in Earth orbit or en route to Mars and back. Could they even function in weightlessness? Would be they be zapped by cosmic rays or solar flares? Were meteors a threat to their craft and, if so, how serious a one? With the dawn of the space age after the launch of Sputnik in October, 1957, these data started to trickle in, and they began to inform plans for Mars missions at NASA and elsewhere.

Radiation was much more of a problem than had been anticipated. The discovery of the Van Allen radiation belts around the Earth and measurement of radiation from solar flares and galactic cosmic rays indicated that short voyages were preferable to long ones, and that crews would need shielding from routine radiation and a “storm shelter” during large solar flares. This motivated research into nuclear thermal and ion propulsion systems, which would not only reduce the transit time to and from Mars, but also, being much more fuel efficient than chemical rockets, dramatically reduce the mass of the ships compared to von Braun's flotilla.

Ernst Stuhlinger had been studying electric (ion) propulsion since 1953, and developed a design for constant-thrust, ion powered ships. These were featured in Walt Disney's 1957 program, “Mars and Beyond”, which aired just two months after the launch of Sputnik. This design was further developed by NASA in a 1962 mission design which envisioned five ships with nuclear-electric propulsion, departing for Mars in the early 1980s with a crew of fifteen and cargo and crew landers permitting a one month stay on the red planet. The ships would rotate to provide artificial gravity for the crew on the trip to and from Mars.

In 1965, the arrival of the Mariner 4 spacecraft seemingly drove a stake through the heart of the romantic view of Mars which had persisted since Percival Lowell. Flying by the southern hemisphere of the planet as close as 9600 km, it returned 21 fuzzy pictures which seemed to show Mars as a dead, cratered world resembling the Moon far more than the Earth. There was no evidence of water, nor of life. The atmosphere was determined to be only 1% as dense as that of Earth, not the 10% estimated previously, and composed mostly of carbon dioxide, not nitrogen. With such a thin and hostile atmosphere, there seemed no prospects for advanced life (anything more complicated than bacteria), and all of the ideas for winged Mars landers went away: the martian atmosphere proved just dense enough to pose a problem when slowing down on arrival, but not enough to allow a soft landing with wings or a parachute. The probe had detected more radiation than expected on its way to Mars, indicating crews would need more protection than anticipated, and it showed that robotic probes could do science at Mars without the need to put a crew at risk. I remember staying up and watching these pictures come in (the local television station didn't carry the broadcast, so I watched even more static-filled pictures than the original from a distant station). I can recall thinking, “Well, that's it then. Mars is dead. We'll probably never go there.”

Mars mission planning went on the back burner as the Apollo Moon program went into high gear in the 1960s. Apollo was conceived not as a single-destination project to land on the Moon, but to create the infrastructure for human expansion from the Earth into the solar system, including development of nuclear propulsion and investigation of planetary missions using Apollo derived hardware, mostly for flyby missions. In January of 1968, Boeing completed a study of a Mars landing mission, which would have required six launches of an uprated Saturn V, sending a crew of six to Mars in a 140 ton ship for a landing and a brief “flags and footprints” stay on Mars. By then, Apollo funding (even before the first lunar orbit and landing) was winding down, and it was clear there was no budget nor political support for such grandiose plans.

After the success of Apollo 11, NASA retrenched, reducing its ambition to a Space Shuttle. An ambitious Space Task Group plan for using the Shuttle to launch a Mars mission in the early 1980s was developed, but in an era of shrinking budgets and additional fly-by missions returning images of a Moon-like Mars, went nowhere. The Saturn V and the nuclear rocket which could have taken crews to Mars had been cancelled. It appeared the U.S. would remain stuck going around in circles in low Earth orbit. And so it remains today.

While planning for manned Mars missions stagnated, the 1970s dramatically changed the view of Mars. In 1971, Mariner 9 went into orbit around Mars and returned 7329 sharp images which showed the planet to be a complex world, with very different northern and southern hemispheres, a grand canyon almost as long as the United States, and features which suggested the existence, at least in the past, of liquid water. In 1976, two Viking orbiters and landers arrived at Mars, providing detailed imagery of the planet and ground truth. The landers were equipped with instruments intended to detect evidence of life, and they reported positive results, but later analyses attributed this to unusual soil chemistry. This conclusion is still disputed, including by the principal investigator for the experiment, but in any case the Viking results revealed a much more complicated and interesting planet than had been imagined from earlier missions. I had been working as a consultant at the Jet Propulsion Laboratory during the first Viking landing, helping to keep mission critical mainframe computers running, and I had the privilege of watching the first images from the surface of Mars arrive. I revised my view from 1965: now Mars was a place which didn't look much different from the high desert of California, where you could imagine going to explore and live some day. More importantly, detailed information about the atmosphere and surface of Mars was now in hand, so future missions could be planned accordingly.

And then…nothing. It was a time of malaise and retreat. After the last Viking landing in September of 1975, it would be more than twenty-one years until Mars Global Surveyor would orbit Mars and Mars Pathfinder would land there in 1996. And yet, with detailed information about Mars in hand, the intervening years were a time of great ferment in manned Mars mission planning, when the foundation of what may be the next great expansion of the human presence into the solar system was laid down.

President George H. W. Bush announced the Space Exploration Initiative on July 20th, 1989, the 20th anniversary of the Apollo 11 landing on the Moon. This was, in retrospect, the last gasp of the “Battlestar” concepts of missions to Mars. It became a bucket into which every NASA centre and national laboratory could throw their wish list: new heavy launchers, a Moon base, nuclear propulsion, space habitats: for a total price tag on the order of half a trillion dollars. It died, quietly, in congress.

But the focus was moving from leviathan bureaucracies of the coercive state to innovators in the private sector. In the 1990s, spurred by work of members of the “Mars Underground”, including Robert Zubrin and David Baker, the “Mars Direct” mission concept emerged. Earlier Mars missions assumed that all resources needed for the mission would have to be launched from Earth. But Zubrin and Baker realised that the martian atmosphere, based upon what we had learned from the Viking missions, contained everything needed to provide breathable air for the stay on Mars and rocket fuel for the return mission (with the addition of lightweight hydrogen brought from Earth). This turned the weight budget of a Mars mission upside-down. Now, an Earth return vehicle could be launched to Mars with empty propellant tanks. Upon arrival, it would produce fuel for the return mission and oxygen for the crew. After it was confirmed to have produced the necessary consumables, the crew of four would be sent in the next launch window (around 26 months later) and land near the return vehicle. They would use its oxygen while on the planet, and its fuel to return to Earth at the end of its mission. There would be no need for a space station in Earth orbit, nor orbital assembly, nor for nuclear propulsion: the whole mission could be done with hardware derived from that already in existence.

This would get humans to Mars, but it ran into institutional barriers at NASA, since many of its pet projects, including the International Space Station and Space Shuttle proved utterly unnecessary to getting to Mars. NASA responded with the Mars Design Reference Mission, published in various revisions between 1993 and 2014, which was largely based upon Mars Direct, but up-sized to a larger crew of six, and incorporating a new Earth Return Vehicle to bring the crew back to Earth in less austere circumstances than envisioned in Mars Direct.

NASA claim they are on a #JourneyToMars. They must be: there's a Twitter hashtag! But of course to anybody who reads this sad chronicle of government planning for planetary exploration over half a century, it's obvious they're on no such thing. If they were truly on a journey to Mars, they would be studying and building the infrastructure to get there using technologies such as propellant depots and in-orbit assembly which would get the missions done economically using resources already at hand. Instead, it's all about building a huge rocket which will cost so much it will fly every other year, at best, employing a standing army which will not only be costly but so infrequently used in launch operations they won't have the experience to operate the system safely, and whose costs will vacuum out the funds which might have been used to build payloads which would extend the human presence into space.

The lesson of this is that when the first humans set foot upon Mars, they will not be civil servants funded by taxes paid by cab drivers and hairdressers, but employees (and/or shareholders) of a private venture that sees Mars as a profit centre which, as its potential is developed, can enrich them beyond the dreams of avarice and provide a backup for human civilisation. I trust that when the history of that great event is written, it will not be as exasperating to read as this chronicle of the dead-end of government space programs making futile efforts to get to Mars.

This is an excellent history of the first half century of manned Mars mission planning. Although many proposed missions are omitted or discussed only briefly, the evolution of mission plans with knowledge of the destination and development of spaceflight hardware is described in detail, culminating with current NASA thinking about how best to accomplish such a mission. This book was published in 2001, but since existing NASA concepts for manned Mars missions are still largely based upon the Design Reference Mission described here, little has changed in the intervening fifteen years. In September of 2016, SpaceX plans to reveal its concepts for manned Mars missions, so we'll have to wait for the details to see how they envision doing it.

As a NASA publication, this book is in the public domain. The book can be downloaded for free as a PDF file from the NASA History Division. There is a paperback republication of this book available at Amazon, but at an outrageous price for such a short public domain work. If you require a paper copy, it's probably cheaper to download the PDF and print your own.

June 2016 Permalink

Pournelle, Jerry. A Step Farther Out. Studio City, CA: Chaos Manor Press, [1979, 1994] 2011. ASIN B004XTKFWW.
This book is a collection of essays originally published in Galaxy magazine between 1974 and 1978. They were originally collected into a book published in 1979, which was republished in 1994 with a new preface and notes from the author. This electronic edition includes all the material from the 1994 book plus a new preface which places the essays in the context of their time and the contemporary world.

I suspect that many readers of these remarks may be inclined to exclaim “Whatever possessed you to read a bunch of thirty-year-old columns from a science fiction magazine which itself disappeared from the scene in 1980?” I reply, “Because the wisdom in these explorations of science, technology, and the human prospect is just as relevant today as it was when I first read them in the original book, and taken together they limn the lost three decades of technological progress which have so blighted our lives.” Pournelle not only envisioned what was possible as humanity expanded its horizons from the Earth to become a spacefaring species drawing upon the resources of the solar system which dwarf those about which the “only one Earth” crowd fret, he also foresaw the constraint which would prevent us from today living in a perfectly achievable world, starting from the 1970s, with fusion, space power satellites, ocean thermal energy conversion, and innovative sources of natural gas providing energy; a robust private space infrastructure with low cost transport to Earth orbit; settlements on the Moon and Mars; exploration of the asteroids with an aim to exploit their resources; and compounded growth of technology which would not only permit human survival but “survival with style”—not only for those in the developed countries, but for all the ten billion who will inhabit this planet by the middle of the present century.

What could possibly go wrong? Well, Pournelle nails that as well. Recall whilst reading the following paragraph that it was written in 1978.

[…] Merely continue as we are now: innovative technology discouraged by taxes, environmental impact statements, reports, lawsuits, commission hearings, delays, delays, delays; space research not carried out, never officially abandoned but delayed, stretched-out, budgets cut and work confined to the studies without hardware; solving the energy crisis by conservation, with fusion research cut to the bone and beyond, continued at level-of-effort but never to a practical reactor; fission plants never officially banned, but no provision made for waste disposal or storage so that no new plants are built and the operating plants slowly are phased out; riots at nuclear power plant construction sites; legal hearings, lawyers, lawyers, lawyers…

Can you not imagine the dream being lost? Can you not imagine the nation slowly learning to “do without”, making “Smaller is Better” the national slogan, fussing over insulating attics and devoting attention to windmills; production falling, standards of living falling, until one day we discover the investments needed to go to space would be truly costly, would require cuts in essentials like food —

A world slowly settling into satisfaction with less, until there are no resources to invest in That Buck Rogers Stuff?

I can imagine that.

As can we all, as now we are living it. And yet, and yet…. One consequence of the Three Lost Decades is that the technological vision and optimistic roadmap of the future presented in these essays is just as relevant to our predicament today as when they were originally published, simply because with a few exceptions we haven't done a thing to achieve them. Indeed, today we have fewer resources with which to pursue them, having squandered our patrimony on consumption, armies of rent-seekers, and placed generations yet unborn in debt to fund our avarice. But for those who look beyond the noise of the headlines and the platitudes of politicians whose time horizon is limited to the next election, here is a roadmap for a true step farther out, in which the problems we perceive as intractable are not “managed” or “coped with”, but rather solved, just as free people have always done when unconstrained to apply their intellect, passion, and resources toward making their fortunes and, incidentally, creating wealth for all.

This book is available only in electronic form for the Kindle as cited above, under the given ASIN. The ISBN of the original 1979 paperback edition is 978-0-441-78584-1. The formatting in the Kindle edition is imperfect, but entirely readable. As is often the case with Kindle documents, “images and tables hardest hit”: some of the tables take a bit of head-scratching to figure out, as the Kindle (or at least the iPad application which I use) particularly mangles multi-column tables. (I mean, what's with that, anyway? LaTeX got this perfectly right thirty years ago, and in a manner even beginners could use; and this was pure public domain software anybody could adopt. Sigh—three lost decades….) Formatting quibbles aside, I'm as glad I bought and read this book as I was when I first bought it and read it all those years ago. If you want to experience not just what the future could have been, then, but what it can be, now, here is an excellent place to start.

The author's Web site is an essential resource for those interested in these big ideas, grand ambitions, and the destiny of humankind and its descendents.

June 2012 Permalink

Powell, James, George Maise, and Charles Pellegrino. StarTram. Seattle: CreateSpace, 2013. ISBN 978-1-4935-7757-6.
Magnetic levitation allows suspending a vehicle above a guideway by the force of magnetic repulsion. A train using magnetic levitation avoids the vibration, noise, and rolling resistance of wheels on rails, and its speed is limited only by air resistance and the amount of acceleration passengers consider tolerable. The Shanghai Maglev Train, in service since 2004, is the fastest train in commercial passenger service today, and travels at 431 kilometres per hour in regular operation. Suppose you were able to somehow get rid of the air resistance and carry only cargo, which can tolerate high acceleration. It would appear that if the technical challenges could be met, the sky would be the limit. In this book the authors argue that the sky is just the start.

They propose a space launch system called StarTram, to be developed in two technological generations. The Generation 1 (Gen-1) system is for cargo only, and uses an evacuated launch tube 110 km long in an underground tunnel. This sounds ambitious, but the three tunnels under the English Channel total 150 km, and are much larger than that required for StarTram. The launcher will be located at a site which allows the tube to run up a mountain, emerging in the thinner air at an altitude between 3 and 7 kilometres. There will be an extreme sonic boom as the launch vehicle emerges from the launch tube at a velocity of around 8 kilometres per second and flies upward through the atmosphere, so the launcher will have to be located in a region where the trajectory downrange for a sufficient distance is unpopulated. Several candidate sites on different continents are proposed.

The Gen-1 cargo craft is levitated by means of high (liquid nitrogen) temperature superconducting magnets which are chilled immediately before launch. They need only remain superconducting for the launch itself, around 30 seconds, so a small on-board supply of liquid nitrogen will suffice for refrigeration. These superconducting magnets repel loops of aluminium in the evacuated guideway tube; no refrigeration of these loops is required. One of the greatest technical challenges of the system is delivering the electric power needed to accelerate the cargo craft. In the 30 seconds or so of acceleration at 30 gravities, the average power requirement is 47 gigawatts, with a peak of 94 gigawatts as orbital velocity is approached. A typical commercial grid power plant produces around 1 gigawatt of power, so it is utterly impractical to generate this power on site. But the total energy required for a launch is only about 20 minutes' output from a 1 gigawatt power station. The StarTram design, therefore, incorporates sixty superconducting energy storage loops, which accumulate the energy for a launch from the grid over time, then discharge to propel the vehicle as it is accelerated. The authors note that the energy storage loops are comparable in magnitude to the superconducting magnets of the Large Hadron Collider, and require neither the extreme precision nor the liquid helium refrigeration those magnets do.

You wouldn't want to ride a Gen-1 cargo launcher. It accelerates at around 30 gravities as it goes down the launch tube, then when it emerges into the atmosphere, decelerates at a rate between 6 and 12g until it flies into the thinner atmosphere. Upon reaching orbital altitude, a small rocket kick motor circularises the orbit. After delivering the payload into orbit (if launching to a higher orbit or one with a different inclination, the payload would contain its own rocket or electric propulsion to reach the desired orbit), the cargo vehicle would make a deorbit burn with the same small rocket it used to circularise its orbit, extend wings, and glide back for re-use.

You may be wondering how a tunnel, evacuated to a sufficiently low pressure to allow a craft to accelerate to orbital velocity without being incinerated, works exactly when one end has to be open to allow the vehicle to emerge into the atmosphere. That bothers me too, a lot. The authors propose that the exit end of the tube will have a door which pops open just before the vehicle is about to emerge. The air at the exit will be ionised by seeding with a conductive material, such as cæsium vapour, then pumped outward by a strong DC current, operating as the inverse of a magnetohydrodynamic generator. Steam generators at the exit of the launch tube force away the ambient air, reducing air pressure as is done for testing upper stage rocket motors. This is something I'd definitely want to see prototyped in both small and full scale before proceeding. Once the cargo craft has emerged, the lid slams shut.

Launching 10 cargo ships a day, the Gen-1 system could deliver 128,000 tons of payload into orbit a year, around 500 times that of all existing rocket launch systems combined. The construction cost of the Gen-1 system is estimated at around US$20 billion, and with all major components reusable, its operating cost is electricity, maintenance, staff, and the small amount of rocket fuel expended in circularising the orbit of craft and deorbiting them. The estimated all-up cost of launching a kilogram of payload is US$43, which is about one hundredth of current launch costs. The launch capacity is adequate to build a robust industrial presence in space, including solar power satellites which beam power to the Earth.

Twenty billion dollars isn't small change, but it's comparable to the development budget for NASA's grotesque Space Launch System, which will fly only every few years and cost on the order of US$2 billion per launch, with everything being thrown away on each mission.

As noted, the Gen-1 system is unsuited to launching people. You could launch people in it, but they wouldn't still be people when they arrived on orbit, due to the accelerations experienced. To launch people, a far more ambitious Gen-2 system is proposed. To reduce launch acceleration to acceptable levels, the launch tunnel would have to be around 1500 km long. To put this into perspective, that's about the distance from Los Angeles to Seattle. To avoid the bruising deceleration (and concomitant loss of velocity) when the vehicle emerges from the launch tube, the end of the launch tube will be magnetically levitated by superconducting magnets (restrained by tethers) so that the end is at an altitude of 20 km. Clearly there'll have to be a no-fly zone around the levitated launch tube, and you really don't want the levitation system to fail. The authors estimate the capital cost of the Gen-2 system at US$67 billion, which seems wildly optimistic to me. Imagine how many forms you'll have to fill out to dig a 1500 km tunnel anywhere in the world, not to speak of actually building one, and then you have to develop that massive magnetically levitated launch tube, which has never been demonstrated.

Essentially everything I have described so far appears in chapter 2 of this book, which makes up less than 10% of its 204 pages. You can read a complete description of the StarTram system for free in this technical paper from 2010. The rest of the book is, well, a mess. With its topic, magnetic levitation space launch, dispensed with by the second chapter, it then veers into describing all of the aspects of our bright future in space such a system will open, including solar power satellites, protecting the Earth from asteroid and comet impacts, space tourism, colonising Mars, exploring the atmosphere of Jupiter, searching for life on the moons of the outer planets, harvesting helium-3 from the atmospheres of the outer planets for fusion power, building a telescope at the gravitational lensing point of the Sun, and interstellar missions. Dark scenarios are presented in which the country which builds StarTram first uses it to establish a global hegemony enforced by all-seeing surveillance from space and “Rods from God”, orbited in their multitudes by StarTram, and a world where the emerging empire is denied access to space by a deliberate effort by one or more second movers to orbit debris to make any use of low orbits impossible, imprisoning humanity on this planet. (But for how long? Small particles in low orbit decay pretty quickly.) Even wilder speculations about intelligent life in the universe and an appropriate strategy for humans in the face of a potentially hostile universe close the book.

All of this is fine, but none of it is new. The only new concept here is StarTram itself, and if the book concentrated just on that, it would be a mere 16 pages. The rest is essentially filler, rehashing other aspects of the human future in space, which would be enabled by any means of providing cheap access to low Earth orbit. The essential question is whether the key enabling technologies of StarTram will work, and that is a matter of engineering which can be determined by component tests before committing to the full-scale project. Were I the NASA administrator and had the power to do so (which, in reality, the NASA administrator does not, being subordinate to the will of appropriators in Congress who mandate NASA priorities in the interest of civil service and contractor jobs in their districts and states), I would cancel the Space Launch System in an instant and use a small part of the savings to fund risk reduction and component tests of the difficult parts of a Gen-1 StarTram launcher.

July 2015 Permalink

Roach, Mary. Packing for Mars. New York: W. W. Norton, 2010. ISBN 978-0-393-06847-4.
At the dawn of the space age, nobody had any idea what effects travel into space might have on living beings, foremost among them the intrepid pilots of the first ships to explore the void. No organism from the ancestral cell of all terrestrial life up to the pointiest-headed professor speculating about its consequences had ever experienced more than an instant of weightlessness, and that usually ended badly with a sudden stop against an unyielding surface. (Fish and human divers are supported by their buoyancy in the water, but they are not weightless: the force of Earth's gravity continues to act upon their internal organs, and might prove to be essential for their correct functioning.) The eye, for example, freed of the pull of gravity, might change shape so that it couldn't focus; it might prove impossible to swallow; digestion of food in the stomach might not work without gravity to hold the contents together at the bottom; urination might fail without gravity working on the contents of the bladder, etc., etc.. The only way to be sure was to go and find out, and this delightful and witty book covers the quest to discover how to live in space, from the earliest animal experiments of the 1940s (most of which ended poorly for the animals, not due to travelling in space, but rather the reliability of the rockets and recovery systems to which they were entrusted) to present day long duration space station missions and research into the human factors of expeditions to Mars and the asteroids.

Travelling to space centres across the U.S., Russia, Europe, and Japan, the author delves into the physiological and psychological, not to mention the humourous and embarrassing aspects of venturing into the vacuum. She boards the vomit comet to experience weightlessness for herself, tries the television camera equipped “aiming practice toilet” on which space shuttle astronauts train before their missions, visits subjects in multi-month bed rest experiments studying loss of muscle and bone mass on simulated interplanetary missions, watches cadavers being used in crash tests of space capsules, tastes a wide variety of overwhelmingly ghastly space food (memo to astronaut corps worldwide: when they hire veterinarians to formulate your chow, don't expect gourmet grub on orbit), and, speaking of grubby, digs into experiments on the outer limits of lack of hygiene, including the odorifically heroic Gemini VII mission in which Frank Borman and James Lovell spent two weeks in a space smaller than the front seat of a Volkswagen Beetle with no way to bathe or open the window, nor bathroom facilities other than plastic bags. Some of the air to ground communications from that mission which weren't broadcast to the public at the time are reproduced here, and are both revealing and amusing in a grody kind of way.

We also meet the animals who preceded the first humans into space, and discover that their personalities were more diverse than those of the Right Stuff humans who followed. You may know of Ham (who was as gung-ho and outgoing as John Glenn) and Enos (who could be as cold and contemptuous as Alan Shepard, and as formidable hurling his feces at those within range as Nolan Ryan was with a baseball), but just imagine those who didn't fly, including Double Ugly, Miss Priss, and Big Mean.

There are a huge number of factoids here, all well-documented, that even the most obsessive space buff may not have come across. For example: why does motion sickness make you vomit? It makes sense to vomit if you've swallowed something truly noxious such as a glass of turpentine or a spoonful of lima beans, but it doesn't make any sense when your visual and vestibular systems are sending conflicting signals since emptying your stomach does nothing to solve the problem. Well, it turns out that functional brain imaging reveals that the “emetic brain” which handles the crucial time-sequencing of the vomit reflex just happens to be located next door in the meat computer to the area which integrates signals from the inner ear and visual system. When the latter is receiving crossed signals, it starts firing neurons wildly trying to make sense of it, and electro-chemical crosstalk gets into vomit central next door and it's a-hurling we will go. It turns out that, despite worries, most human organs work just fine in weightlessness, but some of them behave differently in ways to which space travellers must become accustomed. Consider the bladder—with gravity, the stretching of the wall of the bladder due to the weight of its contents is what triggers the urge to relieve oneself. But in weightlessness, the contents of the bladder, like other fluids, tend to cling to the walls due to surface tension, and the bladder fills up with no signal at all until it's completely full, at which point you have to go right now regardless of whatever you're doing or whether another crewmember is using the space toilet. Reusable manned spacecraft have a certain odour….

There may be nothing that better stimulates the human mind to think out of the box than pondering flight out of this world, and we come across a multitude of examples of innovative boffinology, both from the pages of history and contemporary research. There's the scientist, one of the world's preeminent authorities on chicken brains, who suggested fattening astronauts up to be 20 kilograms obese before launch, which would allow them to fly 90 day missions without the need to launch any food at all. Just imagine the morale among that crew! Not to be outdone, another genius proposed, given the rarity of laundromats in space, that astronauts' clothes be made of digestible fibres, so that they could eat their dirty laundry instead of packaged food. This seems to risk taking “Eat my shorts!” even beyond the tolerance threshold of Bart Simpson. Then consider the people who formulate simulated astronaut poop for testing space toilets, and those who study farts in space. Or, better yet, don't.

If you're remotely interested in space travel, you'll find this a thoroughly enjoyable book, and your only regret when closing it will be that it has come to an end. Speaking of which, if you don't read them as you traverse the main text, be sure to read the extensive end notes—there are additional goodies there for your delectation.

A paperback edition will be published in April 2011.

October 2010 Permalink

Rosen, Milton W. The Viking Rocket Story. New York: Harper & Brothers, 1955. LCCN 55-006592.
This book is out of print. You can generally find used copies at abebooks.com.

February 2003 Permalink

Scott, David and Alexei Leonov with Christine Toomey. Two Sides of the Moon. London: Simon & Schuster, 2004. ISBN 0-7432-3162-7.
Astronaut David Scott flew on the Gemini 8 mission which performed the first docking in space, Apollo 9, the first manned test of the Lunar Module, and commanded the Apollo 15 lunar landing, the first serious scientific exploration of the Moon (earlier Apollo landing missions had far less stay time and, with no lunar rover, limited mobility, and hence were much more “land, grab some rocks, and scoot” exercises). Cosmonaut Alexei Leonov was the first to walk in space on Voskhod 2, led the training of cosmonauts for lunar missions and later the Salyut space station program, and commanded the Soviet side of the Apollo Soyuz Test Project in 1975. Had the Soviet Union won the Moon race, Leonov might well have been first to walk on the Moon. This book recounts the history of the space race as interleaved autobiographies of two participants from contending sides, from their training as fighter pilots ready to kill one another in the skies over Europe in the 1950s to Leonov's handshake in space with an Apollo crew in 1975. This juxtaposition works very well, and writer Christine Toomey (you're not a “ghostwriter” when your name appears on the title page and the principals effusively praise your efforts) does a marvelous job in preserving the engaging conversational style of a one-on-one interview, which is even more an achievement when one considers that she interviewed Leonov through an interpreter, then wrote his contributions in English which was translated to Russian for Leonov's review, with his comments in Russian translated back to English for incorporation in the text. A U.S. edition is scheduled for publication in October 2004.

August 2004 Permalink

Shayler, David J. Apollo: The Lost and Forgotten Missions. London: Springer-Praxis, 2002. ISBN 1-85233-575-0.
Space history buffs will find this well-documented volume fully as fascinating as the title suggests. For a Springer publication, there are a dismaying number of copyediting errors, but I noted only a few errors of fact.

November 2002 Permalink

Siddiqi, Asif A. Challenge to Apollo. Washington: National Aeronautics and Space Administration, 2000. NASA SP-2000-4408.
Prior to the collapse of the Soviet Union, accounts of the Soviet space program were a mix of legend, propaganda, speculations by Western analysts, all based upon a scanty collection of documented facts. The 1990s saw a wealth of previously secret information come to light (although many primary sources remain unavailable), making it possible for the first time to write an authoritative scholarly history of Soviet space exploration from the end of World War II through the mid-1970s; this book, published by the NASA History Division in 2000, is that history.

Whew! Many readers are likely to find that reading this massive (1011 7×14 cm pages, 1.9 kg) book cover to cover tells them far, far more about the Soviet space effort than they ever wanted to know. I bought the book from the U.S. Government Printing Office when it was published in 2000 and have been using it as a reference since then, but decided finally, as the bloggers say, to “read the whole thing”. It was a chore (it took me almost three weeks to chew through it), but ultimately rewarding and enlightening.

Back in the 1960s, when observers in the West pointed out the failure of the communist system to feed its own people or provide them with the most basic necessities, apologists would point to the successes of the Soviet space program as evidence that central planning and national mobilisation in a military-like fashion could accomplish great tasks more efficiently than the chaotic, consumer-driven market economies of the West. Indeed, with the first satellite, the first man in space, long duration piloted flights, two simultaneous piloted missions, the first spacecraft with a crew of more than one, and the first spacewalk, the Soviets racked up an impressive list of firsts. The achievements were real, but based upon what we now know from documents released in the post-Soviet era which form the foundation of this history, the interpretation of these events in the West was a stunning propaganda success by the Soviet Union backed by remarkably little substance.

Indeed, in the 1945–1974 time period covered here, one might almost say that the Soviet Union never actually had a space program at all, in the sense one uses those words to describe the contemporary activities of NASA. The early Soviet space achievements were all spin-offs of ballistic missile technology driven by Army artillery officers become rocket men. Space projects, and especially piloted flight, interested the military very little, and the space spectaculars were sold to senior political figures for their propaganda value, especially after the unanticipated impact of Sputnik on world opinion. But there was never a roadmap for the progressive development of space capability, such as NASA had for projects Mercury, Gemini, and Apollo. Instead, in most cases, it was only after a public success that designers and politicians would begin to think of what they could do next to top that.

Not only did this supposedly centrally planned economy not have a plan, the execution of its space projects was anything but centralised. Throughout the 1960s, there were constant battles among independent design bureaux run by autocratic chief designers, each angling for political support and funding at the expense of the others. The absurdity of this is perhaps best illustrated by the fact that on November 17th, 1967, six days after the first flight of NASA's Saturn V, the Central Committee issued a decree giving the go-ahead to the Chelomey design bureau to develop the UR-700 booster and LK-700 lunar spacecraft to land two cosmonauts on the Moon, notwithstanding having already spent millions of rubles on Korolev's already-underway N1-L3 project, which had not yet performed its first test flight. Thus, while NASA was checking off items in its Apollo schedule, developed years before, the Soviet Union, spending less than half of NASA's budget, found itself committed to two completely independent and incompatible lunar landing programs, with a piloted circumlunar project based on still different hardware simultaneously under development (p. 645).

The catastrophes which ensued from this chaotic situation are well documented, as well as how effective the Soviets were in concealing all of this from analysts in the West. Numerous “out there” proposed projects are described, including Chelomey's monster UR-700M booster (45 million pounds of liftoff thrust, compared to 7.5 million for the Saturn V), which would send a crew of two cosmonauts on a two-year flyby of Mars in an MK-700 spacecraft with a single launch. The little-known Soviet spaceplane projects are documented in detail.

This book is written in the same style as NASA's own institutional histories, which is to say that much of it is heroically boring and dry as the lunar regolith. Unless you're really into reorganisations, priority shifts, power grabs, and other manifestations of gigantic bureaucracies doing what they do best, you may find this tedious. This is not the fault of the author, but of the material he so assiduously presents. Regrettably, the text is set in a light sans-serif font in which (at least to my eyes) the letter “l” and the digit “1” are indistinguishable, and differ from the letter “I” in a detail I can spot only with a magnifier. This, in a book bristling with near-meaningless Soviet institutional names such as the Ministry of General Machine Building and impenetrable acronyms such as NII-1, TsKBEM (not to be confused with TsKBM) and 11F615, only compounds the reader's confusion. There are a few typographical errors, but none are serious.

This NASA publication was never assigned an ISBN, so looking it up on online booksellers will generally only find used copies. You can order new copies from the NASA Information Center at US$79 each. As with all NASA publications, the work is in the public domain, and a scanned online edition (PDF) is available. This is a 64 megabyte download, so unless you have a fast Internet connection, you'll need to be patient. Be sure to download it to a local file as opposed to viewing it in your browser, because otherwise you'll have to download the whole thing each time you open the document.

April 2008 Permalink

Simberg, Rand. Safe Is Not an Option. Jackson, WY: Interglobal Media, 2013. ISBN 978-0-9891355-1-1.
On August 24th, 2011 the third stage of the Soyuz-U rocket carrying the Progress M-12M cargo craft to the International Space Station (ISS) failed during its burn, causing the craft and booster to fall to Earth in Russia. While the crew of six on board the ISS had no urgent need of the supplies on board the Progress, the booster which had failed launching it was essentially identical to that which launched crews to the station in Soyuz spacecraft. Until the cause of the failure was determined and corrected, the launch of the next crew of three, planned for a few weeks later, would have to be delayed. With the Space Shuttle having been retired after its last mission in July 2011, the Soyuz was the only way for crews to reach or return from the ISS. Difficult decisions had to be made, since Soyuz spacecraft in orbit are wasting assets.

The Soyuz has a guaranteed life on orbit of seven months. Regular crew rotations ensure the returning crew does not exceed this “use before” date. But with the launch of new Soyuz missions delayed, it was possible that three crew members would have to return in October before their replacements could arrive in a new Soyuz, and that the remaining three would be forced to leave as well before their craft expired in January. An extended delay while the Soyuz booster problem was resolved would force ISS managers to choose between leaving a skeleton crew of three on board without a known to be safe lifeboat or abandoning the ISS, running the risk that the station, which requires extensive ongoing maintenance by the crew and had a total investment through 2010 estimated at US$ 150 billion might be lost. This was seriously considered.

Just how crazy are these people? The Amundsen-Scott Station at the South Pole has an over-winter crew of around 45 people and there is no lifeboat attached which will enable them, in case of disaster, to be evacuated. In case of fire (considered the greatest risk), the likelihood of mounting rescue missions for the entire crew in mid-winter is remote. And yet the station continues to operate, people volunteer to over-winter there, and nobody thinks too much about the risk they take. What is going on here?

It appears that due to a combination of Cold War elevation of astronauts to symbolic figures and the national trauma of disasters such as Apollo I, Challenger, and Columbia, we have come to view these civil servants as “national treasures” (Jerry Pournelle's words from 1992) and not volunteers who do a risky job on a par with test pilots, naval aviators, firemen, and loggers. This, in turn, leads to statements, oft repeated, that “safety is our highest priority”. Well, if that is the case, why fly? Certainly we would lose fewer astronauts if we confined their activities to “public outreach” as opposed to the more dangerous activities in which less exalted personnel engage such as night aircraft carrier landings in pitching deck conditions done simply to maintain proficiency.

The author argues that we are unwilling to risk the lives of astronauts because of a perception that what they are doing, post-Apollo, is not considered important, and it is hard to dispute that assertion. Going around and around in low Earth orbit and constructing a space station whose crew spend most of their time simply keeping it working are hardly inspiring endeavours. We have lost four decades in which the human presence could have expanded into the solar system, provided cheap and abundant solar power from space to the Earth, and made our species multi-planetary. Because these priorities were not deemed important, the government space program's mission was creating jobs in the districts of those politicians who funded it, and it achieved that.

After reviewing the cost in human life of the development of various means of transportation and exploring our planet, the author argues that we need to be realistic about the risks assumed by those who undertake the task of moving our species off-planet and acknowledge that some of them will not come back, as has been the case in every expansion of the biosphere since the first creature ventured for a brief mission from its home in the sea onto the hostile land. This is not to say that we should design our vehicles and missions to kill their passengers: as we move increasingly from coercively funded government programs to commercial ventures the maxim (too obvious to figure in the Ferengi Rules of Acquisition) “Killing customers is bad for business” comes increasingly into force.

Our focus on “safety first” can lead to perverse choices. Suppose we have a launch system which we estimate that in one in a thousand launches will fail in a way that kills its crew. We equip it with a launch escape system which we estimate that in 90% of the failures will save the crew. So, have we reduced the probability of a loss of crew accident to one in ten thousand? Well, not so fast. What about the possibility that the crew escape mechanism will malfunction and kill the crew on a mission which would have been successful had it not been present? What if solid rockets in the crew escape system accidentally fire in the vehicle assembly building killing dozens of workers and destroying costly and difficult to replace infrastructure? Doing a total risk assessment of such matters is difficult and one gets the sense that little of this is, or will, be done while “safety is our highest priority” remains the mantra.

There is a survey of current NASA projects, including the grotesque “Space Launch System”, a jobs program targeted to the constiuencies of the politicians that mandated it, which has no identified payloads and will be so expensive that it can fly so infrequently the standing army required to maintain it will have little to do between its flights every few years and lose the skills required to operate it safely. Commercial space ventures are surveyed, with a candid analysis of their risks and why the heavy hand of government should allow those willing to accept them to assume them, while protecting the general public from damages from accidents.

The book is superbly produced, with only one typographic error I noted (one “augers” into the ground, nor “augurs”) and one awkward wording about the risks of a commercial space vehicle which will be corrected in subsequent editions. There is a list of acronyms and a comprehensive index.

Disclosure: I contributed to the Kickstarter project which funded the publication of this book, and I received a signed copy of it as a reward. I have no financial interest in sales of this book.

February 2014 Permalink

Spufford, Francis. Backroom Boys: The Secret Return of the British Boffin. London: Faber and Faber, 2003. ISBN 0-571-21496-7.
It is rare to encounter a book about technology and technologists which even attempts to delve into the messy real-world arena where science, engineering, entrepreneurship, finance, marketing, and government policy intersect, yet it is there, not solely in the technological domain, that the roots of both great successes and calamitous failures lie. Backroom Boys does just this and pulls it off splendidly, covering projects as disparate as the Black Arrow rocket, Concorde, mid 1980s computer games, mobile telephony, and sequencing the human genome. The discussion on pages 99 and 100 of the dynamics of new product development in the software business is as clear and concise a statement I've seen of the philosophy that's guided my own activities for the past 25 years. While celebrating the technological renaissance of post-industrial Britain, the author retains the characteristic British intellectual's disdain for private enterprise and economic liberty. In chapter 4, he describes Vodaphone's development of the mobile phone market: “It produced a blind, unplanned, self-interested search strategy, capitalism's classic method for exploring a new space in the market where profit may be found.” Well…yes…indeed, but that isn't just “capitalism's” classic method, but the very one employed with great success by life on Earth lo these four and a half billion years (see The Genius Within, April 2003). The wheels fall off in chapter 5. Whatever your position may have been in the battle between Celera and the public Human Genome Project, Spufford's collectivist bias and ignorance of economics (simply correcting the noncontroversial errors in basic economics in this chapter would require more pages than it fills) gets in the way of telling the story of how the human genome came to be sequenced five years before the original estimated date. A truly repugnant passage on page 173 describes “how science should be done”. Taxpayer-funded researchers, a fine summer evening, “floated back downstream carousing, with stubs of candle stuck to the prows, … and the voices calling to and fro across the water as the punts drifted home under the overhanging trees in the green, green, night.“ Back to the taxpayer-funded lab early next morning, to be sure, collecting their taxpayer-funded salaries doing the work they love to advance their careers. Nary a word here of the cab drivers, sales clerks, construction workers and, yes, managers of biotech start-ups, all taxed to fund this scientific utopia, who lack the money and free time to pass their own summer evenings so sublimely. And on the previous page, the number of cells in the adult body of C. elegans is twice given as 550. Gimme a break—everybody knows there are 959 somatic cells in the adult hermaphrodite, 1031 in the male; he's confusing adults with 558-cell newly-hatched L1 larvæ.

May 2004 Permalink

Stafford, Thomas P. with Michael Cassutt. We Have Capture. Washington: Smithsonian Institution Press, 2002. ISBN 1-58834-070-8.

October 2002 Permalink

Stiennon, Patrick J. G., David M. Hoerr, and Doug Birkholz. The Rocket Company. Reston VA, American Institute of Aeronautics and Astronautics, [2005] 2013. ISBN 978-1-56347-696-9.
This is a very curious book. The American Institute of Aeronautics and Astronautics isn't known as a publisher of fiction, and yet here we have, well, not exactly a novel, but something between an insider account of a disruptive technological start-up company along the lines of The Soul of A New Machine and a business school case study of a company which doesn't exist, at least not yet.

John Forsyth, having made a fortune in the computer software industry, decided to invest in what he believed was the next big thing—drastically reducing the cost of access to space and thereby opening a new frontier not just to coercive governments and wealthy tourists but to pioneers willing to bet their future on expanding the human presence beyond the planet. After dropping a tidy sum in a space start-up in the 1990s, he took a step back and looked at what it would take to build a space access business which would have a real probability of being profitable on a time scale acceptable to investors with the resources it would take to fund it.

Having studied a variety of “new space” companies which focussed on providing launch services in competition with incumbent suppliers, he concluded that in the near term reducing the cost of access to orbit would only result in shrinking overall revenue, as demand for launch services was unlikely to expand much even with a substantial reduction in launch cost. But, as he observed, while in the early days of the airline industry most airlines were unprofitable, surviving on government subsidies, aircraft manufacturers such as Boeing did quite well. So, he decided his new venture would be a vendor of spacecraft hardware, leaving operations and sales of launch services to his customers. It's ugly, but it gets you there.

In optimising an aerospace system, you can trade off one property against another. Most present-day launch systems are optimised to provide maximum lift weight to orbit and use expensive lightweight construction and complex, high-performance engines to achieve that goal. Forsyth opted to focus on reusability and launch rate, even at the cost of payload. He also knew that his budget would not permit the development of exotic technologies, so he chose a two stage to orbit design which would use conventional construction techniques and variants of engines with decades of service history.

He also decided that the launcher would be manned. Given the weight of including crew accommodations, an escape system, and life support equipment this might seem an odd decision, but Forsyth envisioned a substantial portion of his initial market to be countries or other groups who wanted the prestige of having their own manned space program and, further, if there was going to be a pilot on board, he or she could handle payload deployment and other tasks which would otherwise require costly and heavy robotics. (I cannot, for the life of me, figure out the rationale for having a pilot in the first stage. Sure, the added weight doesn't hit the payload to orbit as much as in the second stage, but given the very simple trajectory of the first stage the pilot is little more than a passenger.)

The book chronicles the venture from concept, through business plan, wooing of investors, building the engineering team, making difficult design trade-offs, and pitching the new vehicle to potential customers, carefully avoiding the problem of expectations outpacing reality which had been so often the case with earlier commercial space ventures. The text bristles with cost figures and engineering specifications, the latter all in quaint U.S. units including slugs per square foot (ewww…). Chapter 6 includes a deliciously cynical view of systems engineering as performed in legacy aerospace contractors.

I noted several factual and a number of copy-editing errors, but none which call into question the feasibility of the design. The technologies required to make this work are, for the most part, already in existence and demonstrated in other applications, but whether it would be possible to integrate them into a new vehicle with the schedule and budget envisioned here is unclear. I do not understand at all what happens after the orbital stage lands under its parawing. Both the propellant tanks and interstage compartment are “balloon tanks”, stabilised by pressure. This is fine for flight to orbit, orbital operations (where there is no stress on the interstage when it is depressurised for payload deployment), or re-entry, but after the stage lands horizontally how does the pilot exit through the crew hatch without the interstage losing pressure and crumpling on the runway? Some of the plans for lunar and planetary applications in the final few chapters seem wooly to me, but then I haven't seriously thought about what you might do with a reusable launcher with a payload capacity of 2250 kg that can fly once a day.

The illustrations by Doug Birkholz are superb, reminiscent of those by Russell W. Porter in Amateur Telescope Making. Author Stiennon received U.S. patent 5,568,901 in 1996 for a launch system as described in this book.

May 2013 Permalink

Stumpf, David K. Titan II: A History of a Cold War Missile Program. Fayetteville, AR: The University of Arkansas Press, 2000. ISBN 1-55728-601-9.

May 2002 Permalink

Sullivan, Scott P. Virtual Apollo. Burlington, Canada: Apogee Books, 2002. ISBN 1-896522-94-7.
Every time I see an Apollo command module in a museum, I find myself marveling, “How did they cram all that stuff into that tiny little spacecraft?”. Think about it—the Apollo command and service modules provided everything three men needed to spend two weeks in space, navigate autonomously from the Earth to the Moon and back, dock with other spacecraft, enter and leave lunar orbit, re-enter the Earth's atmosphere at interplanetary speed, fly to a precision splash-down, then serve as a boat until the Navy arrived. And if that wasn't enough, most of the subsystems were doubly or triply redundant, so even in the event of failure, the ship could still get the crew back home, which it did on every single flight, even the dicey Apollo 13. And this amazing flying machine was designed on drawing boards in an era before computer-aided interactive solid modeling was even a concept. Virtual Apollo uses computer aided design to help you appreciate the work of genius which was the Apollo spacecraft. The author created more than 200 painstakingly researched and highly detailed solid models of the command and service modules, which were used to produce the renderings in this book. Ever wondered how the Block II outward-opening crew hatch worked? See pages 41–43. How the devil did they make the docking probe removable? Pages 47–49. Regrettably, the attention to detail which went into production of the models and images didn't follow through to the captions and text, which have apparently been spell-checked but never carefully proofread and contain almost a complete set of nerdish stumbles: its/it's, lose/loose, principal/principle, etc. Let's hope these are remedied in a subsequent edition, and especially that the author or somebody equally talented extends this labour of love to include the lunar module as well.

July 2004 Permalink

Sullivan, Scott P. Virtual LM. Burlington, Canada: Apogee Books, 2004. ISBN 1-894959-14-0.
I closed my comments about the author's earlier Virtual Apollo (July 2004) expressing my hope he would extend the project to the Lunar Module (LM). Well, here it is! These books are based on intricate computer solid models created by Sullivan from extensive research, then rendered to show how subsystems fit into the tightly-packed and weight-constrained spacecraft. The differences between the initial “H mission” modules (Apollo 9–14) and the extended stay “J mission” landers of Apollo 15–17 are shown in comparison renderings. In addition, the Lunar Roving Vehicle (moon buggy) used on the J missions is dissected in the same manner as the LM, along with the life support backpack worn by astronauts on the lunar surface. Nothing about the Lunar Module was simple, and no gory detail is overlooked in this book—there are eight pages (40–47) devoted to the door of the scientific equipment bay and the Rube Goldberg-like mechanism used to open it.

Sadly, like Virtual Apollo, this modeling and rendering labour of love is marred by numerous typographical errors in text and captions. From the point where I started counting, I noted 25, which is an unenviable accomplishment in a 250 page book which is mostly pictures. A companion CD-ROM includes the Apollo Operations Handbook, Lunar Module flight documents from Apollo 14–16, and photographs of the LM simulator and test article.

February 2005 Permalink

Ward, Jonathan H. Rocket Ranch. Cham, Switzerland: Springer International, 2015. ISBN 978-3-319-17788-5.
Many books have been written about Project Apollo, with a large number devoted to the lunar and Skylab missions, the Saturn booster rockets which launched them, the Apollo spacecraft, and the people involved in the program. But none of the Apollo missions could have left the Earth without the facilities at the Kennedy Space Center (KSC) in Florida where the launch vehicle and space hardware were integrated, checked out, fuelled, and launched. In many ways, those facilities were more elaborate and complicated than the booster and spacecraft, and were just as essential in achieving the record of success in Saturn and Apollo/Saturn launches. NASA's 1978 official history of KSC Apollo operations, Moonport (available on-line for free), is a highly recommended examination of the design decisions, architecture, management, and operation of the launch site, but it doesn't delve into the nitty-gritty of how the system actually worked.

The present book, subtitled “The Nuts and Bolts of the Apollo Moon Program at Kennedy Space Center” provides that detail. The author's research involved reviewing more than 1200 original documents and interviewing more than 70 people, most veterans of the Apollo era at KSC (many now elderly). One thread that ran through the interviews is that, to a man (and almost all are men), despite what they had done afterward, they recalled their work on Apollo, however exhausting the pace and formidable the challenges, as a high point in their careers. After completing his research, Ward realised he was looking at a 700 page book. His publisher counselled that such a massive tome would be forbidding to many readers. He decided to separate the description of the KSC hardware (this volume) and the operations leading up to a launch (described in the companion title, Countdown to a Moon Launch, which I will review in the future).

The Apollo/Saturn lunar flight vehicle was, at the time, the most complex machine ever built by humans. It contained three rocket stages (all built by different contractors), a control computer, and two separate spacecraft: the command/service modules and lunar module, each of which had their own rocket engines, control thrusters, guidance computers, and life support systems for the crew. From the moment this “stack” left the ground, everything had to work. While there were redundant systems in case of some in-flight failures, loss of any major component would mean the mission would be unsuccessful, even if the crew returned safely to Earth.

In order to guarantee this success, every component in the booster and spacecraft had to be tested and re-tested, from the time it arrived at KSC until the final countdown and launch. Nothing could be overlooked, and there were written procedures which were followed for everything, with documentation of each step and quality inspectors overseeing it all. The volume of paperwork was monumental (a common joke at the time was that no mission could launch until the paperwork weighed more than the vehicle on the launch pad), but the sheer complexity exceeded the capabilities of even the massive workforce and unlimited budget of Project Apollo. KSC responded by pioneering the use of computers to check out the spacecraft and launcher at every step in the assembly and launch process. Although a breakthrough at the time, the capacity of these computers is laughable today. The computer used to check out the Apollo spacecraft had 24,576 words of memory when it was installed in 1964, and programmers had to jump through hoops and resort to ever more clever tricks to shoehorn the test procedures into the limited memory. Eventually, after two years, approval was obtained to buy an additional 24,000 words of memory for the test computers, at a cost of almost half a million 2015 dollars.

You've probably seen pictures of the KSC firing room during Apollo countdowns. The launch director looked out over a sea of around 450 consoles, each devoted to one aspect of the vehicle (for example, console BA25, “Second stage propellant utilization”), each manned by an engineer in a white shirt and narrow tie. These consoles were connected into audio “nets”, arranged in a hierarchy paralleling the management structure. For example, if the engineer at console BA25 observed something outside acceptable limits, he would report it on the second stage propulsion net. The second stage manager would then raise the issue on the launch vehicle net. If it was a no-go item, it would then be bumped up to the flight director loop where a hold would be placed on the countdown. If this wasn't complicated enough, most critical parameters were monitored by launch vehicle and spacecraft checkout computers, which could automatically halt the countdown if a parameter exceeded limits. Most of those hundreds of consoles had dozens of switches, indicator lights, meters, and sometimes video displays, and all of them had to be individually wired to patchboards which connected them to the control computers or, in some cases, directly to the launch hardware. And every one of those wires had to have a pull ticket for its installation, and inspection, and an individual test and re-test that it was functioning properly. Oh, and there were three firing rooms, identically equipped. During a launch, two would be active and staffed: one as a primary, the other as a backup.

The level of detail here is just fantastic and may be overwhelming if not taken in small doses. Did you know, for example, that in the base of the Saturn V launch platform there was an air conditioned room with the RCA 110A computer which checked out the booster? The Saturn V first stage engines were about 30 metres from this delicate machine. How did they keep it from being pulverised when the rocket lifted off? Springs.

Assembled vehicles were transported from the Vehicle Assembly Building to the launch pad by an enormous crawler. The crawler was operated by a crew of 14, including firemen stationed near the diesel engines. Originally, there was an automatic fire suppression system, but after it accidentally triggered and dumped a quarter ton of fire suppression powder into one of the engines during a test, it was replaced with firemen. How did they keep the launcher level as it climbed up the ramp to the pad? They had two pipes filled with mercury which ran diagonally across the crawler platform between each pair of corners. These connected to a sight glass which indicated to the operator if the platform wasn't level. Then the operator would adjust jacking cylinders on the corners to restore the platform to level—while it was rolling.

I can provide only a few glimpses of the wealth of fascinating minutæ on all aspects of KSC facilities and operations described here. Drawing on his more than 300 hours of interviews, the author frequently allows veterans of the program to speak in their own words, giving a sense of what it was like to be there, then, the rationale for why things were done the way they were, and to relate anecdotes about when things didn't go as planned.

It has been said that one of the most difficult things NASA did in Project Apollo was to make it look easy. Even space buffs who have devoured dozens of books about Apollo may be startled by the sheer magnitude of what was accomplished in designing, building, checking out, and operating the KSC facilities described in this book, especially considering in how few years it all was done and the primitive state of some of the technologies available at the time (particularly computers and electronics). This book and its companion volume are eye-openers, and only reinforce what a technological triumph Apollo was.

December 2015 Permalink

Ward, Jonathan H. Countdown to a Moon Launch. Cham, Switzerland: Springer International, 2015. ISBN 978-3-319-17791-5.
In the companion volume, Rocket Ranch (December 2015), the author describes the gargantuan and extraordinarily complex infrastructure which was built at the Kennedy Space Center (KSC) in Florida to assemble, check out, and launch the Apollo missions to the Moon and the Skylab space station. The present book explores how that hardware was actually used, following the “processing flow” of the Apollo 11 launch vehicle and spacecraft from the arrival of components at KSC to the moment of launch.

As intricate as the hardware was, it wouldn't have worked, nor would it have been possible to launch flawless mission after flawless mission on time had it not been for the management tools employed to coordinate every detail of processing. Central to this was PERT (Program Evaluation and Review Technique), a methodology developed by the U.S. Navy in the 1950s to manage the Polaris submarine and missile systems. PERT breaks down the progress of a project into milestones connected by activities into a graph of dependencies. Each activity has an estimated time to completion. A milestone might be, say, the installation of the guidance system into a launch vehicle. That milestone would depend upon the assembly of the components of the guidance system (gyroscopes, sensors, electronics, structure, etc.), each of which would depend upon their own components. Downstream, integrated test of the launch vehicle would depend upon the installation of the guidance system. Many activities proceed in parallel and only come together when a milestone has them as its mutual dependencies. For example, the processing and installation of rocket engines is completely independent of work on the guidance system until they join at a milestone where an engine steering test is performed.

As a project progresses, the time estimates for the various activities will be confronted with reality: some will be completed ahead of schedule while other will slip due to unforeseen problems or over-optimistic initial forecasts. This, in turn, ripples downstream in the dependency graph, changing the time available for activities if the final completion milestone is to be met. For any given graph at a particular time, there will be a critical path of activities where a schedule slip of any one will delay the completion milestone. Each lower level milestone in the graph has its own critical path leading to it. As milestones are completed ahead or behind schedule, the overall critical path will shift. Knowing the critical path allows program managers to concentrate resources on items along the critical path to avoid, wherever possible, overall schedule slips (with the attendant extra costs).

Now all this sounds complicated, and in a project with the scope of Apollo, it is almost bewildering to contemplate. The Launch Control Center was built with four firing rooms. Three were outfitted with all of the consoles to check out and launch a mission, but the fourth cavernous room ended up being used to display and maintain the PERT charts for activities in progress. Three levels of charts were maintained. Level A was used by senior management and contained hundreds of major milestones and activities. Each of these was expanded out into a level B chart which, taken together, tracked in excess of 7000 milestones. These, in turn, were broken down into detail on level C charts, which tracked more than 40,000 activities. The level B and C charts were displayed on more than 400 square metres of wall space in the back room of firing room four. As these detailed milestones were completed on the level C charts, changes would propagate down that chart and those which affected its completion upward to the level A and B charts.

Now, here's the most breathtaking thing about this: they did it all by hand! For most of the Apollo program, computer implementations of PERT were not available (or those that existed could not handle this level of detail). (Today, the PERT network for processing of an Apollo mission could be handled on a laptop computer.) There were dozens of analysts and clerks charged with updating the networks, with the processing flow displayed on an enormous board with magnetic strips which could be shifted around by people climbing up and down rolling staircases. Photographers would take pictures of the board which were printed and distributed to managers monitoring project status.

If PERT was essential to coordinating all of the parallel activities in preparing a spacecraft for launch, configuration control was critical to ensure than when the countdown reached T0, everything would work as expected. Just as there was a network of dependencies in the PERT chart, the individual components were tested, subassemblies were tested, assemblies of them were tested, all leading up to an integrated test of the assembled launcher and spacecraft. The successful completion of a test established a tested configuration for the item. Anything which changed that configuration in any way, for example unplugging a cable and plugging it back in, required re-testing to confirm that the original configuration had been restored. (One of the pins in the connector might not have made contact, for instance.) This was all documented by paperwork signed off by three witnesses. The mountain of paper was intimidating; there was even a slide rule calculator for estimating the cost of various kinds of paperwork.

With all of this management superstructure it may seem a miracle that anything got done at all. But, as the end of the decade approached, the level of activity at KSC was relentless (and took a toll upon the workforce, although many recall it as the most intense and rewarding part of their careers). Several missions were processed in parallel: Apollo 11 rolled out to the launch pad while Apollo 10 was still en route to the Moon, and Apollo 12 was being assembled and tested.

To illustrate how all of these systems and procedures came together, the author takes us through the processing of Apollo 11 in detail, starting around six months before launch when the Saturn V stages, and command, service, and lunar modules arrived independently from the contractors who built them or the NASA facilities where they had been individually tested. The original concept for KSC was that it would be an “operational spaceport” which would assemble pre-tested components into flight vehicles, run integrated system tests, and then launch them in an assembly-line fashion. In reality, the Apollo and Saturn programs never matured to this level, and were essentially development and test projects throughout. Components not only arrived at KSC with “some assembly required”; they often were subject to a blizzard of engineering change orders which required partially disassembling equipment to make modifications, then exhaustive re-tests to verify the previously tested configuration had been restored.

Apollo 11 encountered relatively few problems in processing, so experiences from other missions where problems arose are interleaved to illustrate how KSC coped with contingencies. While Apollo 16 was on the launch pad, a series of mistakes during the testing process damaged a propellant tank in the command module. The only way to repair this was to roll the entire stack back to the Vehicle Assembly Building, remove the command and service modules, return them to the spacecraft servicing building then de-mate them, pull the heat shield from the command module, change out the tank, then put everything back together, re-stack, and roll back to the launch pad. Imagine how many forms had to be filled out. The launch was delayed just one month.

The process of servicing the vehicle on the launch pad is described in detail. Many of the operations, such as filling tanks with toxic hypergolic fuel and oxidiser, which burn on contact, required evacuating the pad of all non-essential personnel and special precautions for those engaged in these hazardous tasks. As launch approached, the hurdles became higher: a Launch Readiness Review and the Countdown Demonstration Test, a full dress rehearsal of the countdown up to the moment before engine start, including fuelling all of the stages of the launch vehicle (and then de-fuelling them after conclusion of the test).

There is a wealth of detail here, including many obscure items I've never encountered before. Consider “Forward Observers”. When the Saturn V launched, most personnel and spectators were kept a safe distance of more than 5 km from the launch pad in case of calamity. But three teams of two volunteers each were stationed at sites just 2 km from the pad. They were charged with observing the first seconds of flight and, if they saw a catastrophic failure (engine explosion or cut-off, hard-over of an engine gimbal, or the rocket veering into the umbilical tower), they would signal the astronauts to fire the launch escape system and abort the mission. If this happened, the observers would then have to dive into crude shelters often frequented by rattlesnakes to ride out the fiery aftermath.

Did you know about the electrical glitch which almost brought the Skylab 2 mission to flaming catastrophe moments after launch? How lapses in handling of equipment and paperwork almost spelled doom for the crew of Apollo 13? The time an oxygen leak while fuelling a Saturn V booster caused cars parked near the launch pad to burst into flames? It's all here, and much more. This is an essential book for those interested in the engineering details of the Apollo project and the management miracles which made its achievements possible.

January 2016 Permalink

Weil, Elizabeth. They All Laughed at Christopher Columbus. New York: Bantam Books, 2002. ISBN 978-0-553-38236-5.
For technologists and entrepreneurs, the latter half of the 1990s was a magical time. The explosive growth in computing power available to individuals, the global interconnectivity afforded by the Internet, and the emergence of broadband service with the potential to make the marginal cost of entry as a radio or video broadcaster next to zero created a vista of boundless technological optimism. Companies with market valuations in the billions sprang up like mushrooms despite having never turned a profit (and in some cases, before delivering a product), and stock-option paper millionaires were everywhere, some sporting job titles which didn't exist three years before.

In this atmosphere enthusiasms of all kinds were difficult to restrain, even those more venerable than Internet start-ups, and among people who had previously been frustrated upon multiple occasions. So it was that as the end of the decade approached, Gary Hudson, veteran of three earlier unsuccessful commercial space projects, founded Rotary Rocket, Inc. with the goal of building a reusable single-stage-to-orbit manned spacecraft which would reduce the cost of launching payloads into low Earth orbit by a factor of ten compared to contemporary expendable rockets (which, in turn, were less expensive than NASA's Space Shuttle). Such a dramatic cost reduction was expected to immediately generate substantial business from customers such as Teledesic, which originally planned to launch 840 satellites to provide global broadband Internet service. Further, at one tenth the launch cost, space applications which were not economically feasible before would become so, expanding the space market just as the comparable collapse in the price of computing and communications had done in their sectors.

Hudson assembled a team, a mix of veterans of his earlier ventures, space enthusiasts hoping to make their dreams a reality at last, hard-nosed engineers, and seasoned test pilots hoping to go to space, and set to work. His vision became known as Roton, and evolved to be an all-composite structure including tanks for the liquid oxygen and kerosene propellants, and a unique rotary engine at the base of the conical structure which would spin to create the pressure to inject propellants into 96 combustors arrayed around the periphery, eliminating the need for heavy, complicated, and prone-to-disintegrate turbopumps. The crew of two would fly the Roton to orbit and release the payload into space, then make a de-orbit burn. During re-entry, a water-cooled heat shield on the base of the cone would protect the structure from heating, and when atmospheric density was sufficient, helicopter-like rotor blades would deploy from the top of the cone. These blades would be spun up by autorotation and then, shortly before touchdown, tip jets powered by hydrogen peroxide would fire to allow a controlled powered approach and precision landing. After a mission, one need only load the next payload, refill the propellant tanks, and brief the crew for the next flight. It was estimated one flight per day was achievable with a total ground staff of fewer than twenty people.

This would have been revolutionary, and there were many, some with forbidding credentials and practical experience, who argued that it couldn't possibly work, and certainly not on Hudson's schedule and budget of US$ 150 million (which is closer to the sum NASA or one of its contractors would require to study such a concept, not to actually build and fly it). There were many things to worry about. Nothing like the rotary engine had ever been built, and its fluid mechanical and thermal complexities were largely unknown. The heat shield was entirely novel, and there was no experience as to how it would perform in a real world environment in which pores and channels might clog. Just getting to orbit in a single stage vehicle powered by LOX and kerosene was considered impossible by many, requiring a structure which was 95% propellant at launch. Even with composite construction, nobody had achieved anything close to this mass fraction in a flight vehicle.

Gary Hudson is not just a great visionary; he is nothing if not persuasive. For example, here is a promotional video from 1998. He was able, over the history of the project, to raise a total of US$ 30 million for the project from private investors (disclosure: myself included), and built an initial atmospheric test vehicle intended to validate the helicopter landing system. In 1999, this vehicle made three successful test flights, including a hop up and down and a flight down the runway.

By this point in 1999, the technology bubble was nearing the bursting point and perspicacious investors were already backing away from risky ventures. When it became clear there was no prospect to raise sufficient funds to continue, even toward the next milestone, Hudson had no option but to lay off staff and eventually entirely shutter the company, selling off its remaining assets (but the Roton ATV can be seen on display at the Mojave Spaceport).

There are any number of “business books” written about successful ventures, often ghostwritten for founders to show how they had a unique vision and marched from success to success to achieve their dream. (These so irritated me that I strove, in my own business book, to demonstrate from contemporary documents, the extent to which those in a technological start-up grope in the dark with insufficient information and little idea of where it's going.) Much rarer are accounts of big dreams which evoked indefatigable efforts from talented people and, despite all, ended badly. This book is a superb exemplar of that rare genre. There are a few errors of fact, and from time to time the author's description of herself among the strange world of the rocket nerds is a bit precious, but you get an excellent sense of what it was like to dream big, how a visionary can inspire people to accomplish extraordinary things, and how an entrepreneur must not only have a sound technical foundation, a vision of the future, but also have kissed the Barnum stone to get the job done.

Oddly, the book contains no photographs of this unique and stunning vehicle or the people who built it.

October 2013 Permalink

Wendt, Guenter and Russell Still. The Unbroken Chain. Burlington, Canada: Apogee Books, 2001. ISBN 1-896522-84-X.

December 2001 Permalink

White, Rowland. Into the Black. New York: Touchstone, 2016. ISBN 978-1-5011-2362-7.
On April 12, 1981, coincidentally exactly twenty years after Yuri Gagarin became the first man to orbit the Earth in Vostok 1, the United States launched one of the most ambitious and risky manned space flights ever attempted. The flight of Space Shuttle Orbiter Columbia on its first mission, STS-1, would be the first time a manned spacecraft was launched with a crew on its first flight. (All earlier spacecraft were tested in unmanned flights before putting a crew at risk.) It would also be the first manned spacecraft to be powered by solid rocket boosters which, once lit, could not be shut down but had to be allowed to burn out. In addition, it would be the first flight test of the new Space Shuttle Main Engines, the most advanced and high performance rocket engines ever built, which had a record of exploding when tested on the ground. The shuttle would be the first space vehicle to fly back from space using wings and control surfaces to steer to a pinpoint landing. Instead of a one-shot ablative heat shield, the shuttle was covered by fragile silica tiles and reinforced carbon-carbon composite to protect its aluminium structure from reentry heating which, without thermal protection, would melt it in seconds. When returning to Earth, the shuttle would have to maneuver in a hypersonic flight regime in which no vehicle had ever flown before, then transition to supersonic and finally subsonic flight before landing. The crew would not control the shuttle directly, but fly it through redundant flight control computers which had never been tested in flight. Although the orbiter was equipped with ejection seats for the first four test flights, they could only be used in a small part of the flight envelope: for most of the mission everything simply had to work correctly for the ship and crew to return safely. Main engine start—ignition of the solid rocket boosters—and liftoff!

Even before the goal of landing on the Moon had been accomplished, it was apparent to NASA management that no national consensus existed to continue funding a manned space program at the level of Apollo. Indeed, in 1966, NASA's budget reached a peak which, as a fraction of the federal budget, has never been equalled. The Saturn V rocket was ideal for lunar landing missions, but expended each mission, was so expensive to build and operate as to be unaffordable for suggested follow-on missions. After building fifteen Saturn V flight vehicles, only thirteen of which ever flew, Saturn V production was curtailed. With the realisation that the “cost is no object” days of Apollo were at an end, NASA turned its priorities to reducing the cost of space flight, and returned to a concept envisioned by Wernher von Braun in the 1950s: a reusable space ship.

You don't have to be a rocket scientist or rocket engineer to appreciate the advantages of reusability. How much would an airline ticket cost if they threw away the airliner at the end of every flight? If space flight could move to an airline model, where after each mission one simply refueled the ship, performed routine maintenance, and flew again, it might be possible to reduce the cost of delivering payload into space by a factor of ten or more. But flying into space is much more difficult than atmospheric flight. With the technologies and fuels available in the 1960s (and today), it appeared next to impossible to build a launcher which could get to orbit with just a single stage (and even if one managed to accomplish it, its payload would be negligible). That meant any practical design would require a large booster stage and a smaller second stage which would go into orbit, perform the mission, then return.

Initial design concepts envisioned a very large (comparable to a Boeing 747) winged booster to which the orbiter would be attached. At launch, the booster would lift itself and the orbiter from the pad and accelerate to a high velocity and altitude where the orbiter would separate and use its own engines and fuel to continue to orbit. After separation, the booster would fire its engines to boost back toward the launch site, where it would glide to a landing on a runway. At the end of its mission, the orbiter would fire its engines to de-orbit, then reenter the atmosphere and glide to a landing. Everything would be reusable. For the next mission, the booster and orbiter would be re-mated, refuelled, and readied for launch.

Such a design had the promise of dramatically reducing costs and increasing flight rate. But it was evident from the start that such a concept would be very expensive to develop. Two separate manned spacecraft would be required, one (the booster) much larger than any built before, and the second (the orbiter) having to operate in space and survive reentry without discarding components. The orbiter's fuel tanks would be bulky, and make it difficult to find room for the payload and, when empty during reentry, hard to reinforce against the stresses they would encounter. Engineers believed all these challenges could be met with an Apollo era budget, but with no prospect of such funds becoming available, a more modest design was the only alternative.

Over a multitude of design iterations, the now-familiar architecture of the space shuttle emerged as the only one which could meet the mission requirements and fit within the schedule and budget constraints. Gone was the flyback booster, and with it full reusability. Two solid rocket boosters would be used instead, jettisoned when they burned out, to parachute into the ocean and be fished out by boats for refurbishment and reuse. The orbiter would not carry the fuel for its main engines. Instead, it was mounted on the side of a large external fuel tank which, upon reaching orbit, would be discarded and burn up in the atmosphere. Only the orbiter, with its crew and payload, would return to Earth for a runway landing. Each mission would require either new or refurbished solid rocket boosters, a new external fuel tank, and the orbiter.

The mission requirements which drove the design were not those NASA would have chosen for the shuttle were the choice theirs alone. The only way NASA could “sell” the shuttle to the president and congress was to present it as a replacement for all existing expendable launch vehicles. That would assure a flight rate sufficient to achieve the economies of scale required to drive down costs and reduce the cost of launch for military and commercial satellite payloads as well as NASA missions. But that meant the shuttle had to accommodate the large and heavy reconnaissance satellites which had been launched on Titan rockets. This required a huge payload bay (15 feet wide by 59 feet long) and a payload to low Earth orbit of 60,000 pounds. Further Air Force requirements dictated a large cross-range (ability to land at destinations far from the orbital ground track), which in turn required a hot and fast reentry very demanding on the thermal protection system.

The shuttle represented, in a way, the unification of NASA with the Air Force's own manned space ambitions. Ever since the start of the space age, the Air Force sought a way to develop its own manned military space capability. Every time it managed to get a program approved: first Dyna-Soar and then the Manned Orbiting Laboratory, budget considerations and Pentagon politics resulted in its cancellation, orphaning a corps of highly-qualified military astronauts with nothing to fly. Many of these pilots would join the NASA astronaut corps in 1969 and become the backbone of the early shuttle program when they finally began to fly more than a decade later.

All seemed well on board. The main engines shut down. The external fuel tank was jettisoned. Columbia was in orbit. Now weightless, commander John Young and pilot Bob Crippen immediately turned to the flight plan, filled with tasks and tests of the orbiter's systems. One of their first jobs was to open the payload bay doors. The shuttle carried no payload on this first flight, but only when the doors were open could the radiators that cooled the shuttle's systems be deployed. Without the radiators, an emergency return to Earth would be required lest electronics be damaged by overheating. The doors and radiators functioned flawlessly, but with the doors open Young and Crippen saw a disturbing sight. Several of the thermal protection tiles on the pods containing the shuttle's maneuvering engines were missing, apparently lost during the ascent to orbit. Those tiles were there for a reason: without them the heat of reentry could melt the aluminium structure they protected, leading to disaster. They reported the missing tiles to mission control, adding that none of the other tiles they could see from windows in the crew compartment appeared to be missing.

The tiles had been a major headache during development of the shuttle. They had to be custom fabricated, carefully applied by hand, and were prone to falling off for no discernible reason. They were extremely fragile, and could even be damaged by raindrops. Over the years, NASA struggled with these problems, patiently finding and testing solutions to each of them. When STS-1 launched, they were confident the tile problems were behind them. What the crew saw when those payload bay doors opened was the last thing NASA wanted to see. A team was set to analysing the consequences of the missing tiles on the engine pods, and quickly reported back that they should pose no problem to a safe return. The pods were protected from the most severe heating during reentry by the belly of the orbiter, and the small number of missing tiles would not affect the aerodynamics of the orbiter in flight.

But if those tiles were missing, mightn't other tiles also have been lost? In particular, what about those tiles on the underside of the orbiter which bore the brunt of the heating? If some of them were missing, the structure of the shuttle might burn through and the vehicle and crew would be lost. There was no way for the crew to inspect the underside of the orbiter. It couldn't be seen from the crew cabin, and there was no way to conduct an EVA to examine it. Might there be other, shall we say, national technical means, of inspecting the shuttle in orbit? Now STS-1 truly ventured into the black, a story never told until many years after the mission and documented thoroughly for a popular audience here for the first time.

In 1981, ground-based surveillance of satellites in orbit was rudimentary. Two Department of Defense facilities, in Hawaii and Florida, normally used to image Soviet and Chinese satellites, were now tasked to try to image Columbia in orbit. This was a daunting task: the shuttle was in a low orbit, which meant waiting until an orbital pass would cause it to pass above one of the telescopes. It would be moving rapidly so there would be only seconds to lock on and track the target. The shuttle would have to be oriented so its belly was aimed toward the telescope. Complicating the problem, the belly tiles were black, so there was little contrast against the black of space. And finally, the weather had to cooperate: without a perfectly clear sky, there was no hope of obtaining a usable image. Several attempts were made, all unsuccessful.

But there were even deeper black assets. The National Reconnaissance Office (whose very existence was a secret at the time) had begun to operate the KH-11 KENNEN digital imaging satellites in the 1970s. Unlike earlier spysats, which exposed film and returned it to the Earth for processing and interpretation, the KH-11 had a digital camera and the ability to transmit imagery to ground stations shortly after it was captured. There were few things so secret in 1981 as the existence and capabilities of the KH-11. Among the people briefed in on this above top secret program were the NASA astronauts who had previously been assigned to the Manned Orbiting Laboratory program which was, in fact, a manned reconnaissance satellite with capabilities comparable to the KH-11.

Dancing around classification, compartmentalisation, bureaucratic silos, need to know, and other barriers, people who understood what was at stake made it happen. The flight plan was rewritten so that Columbia was pointed in the right direction at the right time, the KH-11 was programmed for the extraordinarily difficult task of taking a photo of one satellite from another, when their closing velocities are kilometres per second, relaying the imagery to the ground and getting it to the NASA people who needed it without the months of security clearance that would normally entail. The shuttle was a key national security asset. It was to launch all reconnaissance satellites in the future. Reagan was in the White House. They made it happen. When the time came for Columbia to come home, the very few people who mattered in NASA knew that, however many other things they had to worry about, the tiles on the belly were not among them.

(How different it was in 2003 when the same Columbia suffered a strike on its left wing from foam shed from the external fuel tank. A thoroughly feckless and bureaucratised NASA rejected requests to ask for reconnaissance satellite imagery which, with two decades of technological improvement, would have almost certainly revealed the damage to the leading edge which doomed the orbiter and crew. Their reason: “We can't do anything about it anyway.” This is incorrect. For a fictional account of a rescue, based upon the report [PDF, scroll to page 173] of the Columbia Accident Investigation Board, see Launch on Need [February 2012].)

This is a masterful telling of a gripping story by one of the most accomplished of aerospace journalists. Rowan White is the author of Vulcan 607 (May 2010), the definitive account of the Royal Air Force raid on the airport in the Falkland Islands in 1982. Incorporating extensive interviews with people who were there, then, and sources which were classified until long after the completion of the mission, this is a detailed account of one of the most consequential and least appreciated missions in U.S. manned space history which reads like a techno-thriller.

September 2016 Permalink

Worden, Al with Francis French. Falling to Earth. Washington: Smithsonian Books, 2011. ISBN 978-1-58834-309-3.
Al Worden (his given name is Alfred, but he has gone by “Al” his whole life) was chosen as a NASA astronaut in April 1966, served as backup command module pilot for the Apollo 12 mission, the second Moon landing, and then flew to the Moon as command module pilot of Apollo 15, the first serious geological exploration mission. As command module pilot, Worden did not land on the Moon but, while tending the ship in orbit awaiting the return of his crewmates, operated a series of scientific experiments, some derived from spy satellite technology, which provided detailed maps of the Moon and a survey of its composition. To retrieve the film from the mapping cameras in the service module, Worden performed the first deep-space EVA during the return to Earth.

Growing up on a farm in rural Michigan during the first great depression and the second World War, Worden found his inclination toward being a loner reinforced by the self-reliance his circumstances forced upon him. He remarks on several occasions how he found satisfaction in working by himself and what he achieved on his own and while not disliking the company of others, found no need to validate himself through their opinions of him. This inner-directed drive led him to West Point, which he viewed as the only way to escape from a career on the farm given his family's financial circumstances, an Air Force commission, and graduation from the Empire Test Pilots' School in Farnborough, England under a US/UK exchange program.

For one inclined to be a loner, it would be difficult to imagine a more ideal mission than Worden's on Apollo 15. Orbiting the Moon in the command module Endeavour for almost three days by himself he was, at maximum distance on the far side of the Moon, more isolated from his two crewmates on the surface than any human has been from any other humans before or since (subsequent Apollo missions placed the command module in a lower lunar orbit, reducing this distance slightly). He candidly admits how much he enjoyed being on his own in the capacious command module, half the time entirely his own man while out of radio contact behind the Moon, and how his joy at the successful return of his comrades from the surface was tempered by how crowded and messy the command module was with them, the Moon rocks they collected, and all the grubby Moon dust clinging to their spacesuits on board.

Some Apollo astronauts found it difficult to adapt to life on Earth after their missions. Travelling to the Moon before you turn forty is a particularly extreme case of “peaking early”, and the question of “What next?” can be formidable, especially when the entire enterprise of lunar exploration was being dismantled at its moment of triumph. Still, one should not overstate this point: of the twenty-four astronauts who flew to the Moon, most went on to subsequent careers you'd expect for the kind of overachievers who become astronauts in the first place—in space exploration, the military, business, politics, education, and even fine arts. Few, however, fell to Earth so hard as the crew of Apollo 15. The collapse of one of their three landing parachutes before splashdown due to the canopy's being eroded due to a dump of reaction control propellant might have been seen as a premonition of this, but after the triumphal conclusion of a perfect mission, a White House reception, an address to a joint session of Congress, and adulatory celebrations on a round-the-world tour, it all came undone in an ugly scandal involving, of all things, postage stamps.

The Apollo 15 crew, like those of earlier NASA missions, had carried on board as part of their “personal preference kits” postage stamp covers commemorating the flight. According to Worden's account in this book, the Apollo 15 covers were arranged by mission commander Dave Scott, and agreed to by Worden and lunar module pilot Jim Irwin on Scott's assurance that this was a routine matter which would not affect their careers and that any sales of the covers would occur only after their retirement from NASA and the Air Force (in which all three were officers). When, after the flight, the covers began to come onto the market, an ugly scandal erupted, leading to the Apollo 15 crew being removed from flight status, and Worden and Irwin being fired from NASA with reprimands placed in their Air Force records which would block further promotion. Worden found himself divorced (before the Moon mission), out of a job at NASA, and with no future in the Air Force.

Reading this book, you get the impression that this was something like the end of Worden's life. And yet it wasn't—he went on to complete his career in the flight division at NASA's Ames Research Center and retire with the rank and pension of a Colonel in the U.S. Air Force. He then served in various capacities in private sector aerospace ventures and as chairman of the Astronaut Scholarship Foundation. Honestly, reading this book, you get the sense that everybody has forgotten the stupid postage stamps except the author. If there is some kind of redemption to be had by recounting the episode here (indeed, “Redemption” is the title of chapter 13 of this work), then fine, but whilst reading this account, I found myself inclined to shout, “Dude—you flew to the Moon! Yes, you messed up and got fired—who hasn't? But you landed on your feet and have had a wonderful life since, including thirty years of marriage. Get over the shaggy brown ugliness of the 1970s and enjoy the present and all the years to come!”

October 2011 Permalink

Young, Anthony. The Saturn V F-1 Engine. Chichester, UK: Springer Praxis, 2009. ISBN 978-0-387-09629-2.
The F-1 rocket engine which powered the first (S-IC) stage of the Saturn V booster, which launched all of the Apollo missions to the Moon and, as a two stage variant, the Skylab space station, was one of the singular engineering achievements of the twentieth century, which this magnificent book chronicles in exquisite detail. When the U.S. Air Force contracted with Rocketdyne in 1958 for the preliminary design of a single chamber engine with between 1 and 1.5 million pounds of thrust, the largest existing U.S. rocket engine had less than a quarter the maximum thrust of the proposed new powerplant, and there was no experience base to provide confidence that problems such as ignition transients and combustion instability which bedevil liquid rockets would not prove insuperable when scaling an engine to such a size. (The Soviets were known to have heavy-lift boosters, but at the time nobody knew their engine configuration. In fact, when their details came to be known in the West, they were discovered to use multiple combustion chambers and/or clustering of engines precisely to avoid the challenges of very large engines.)

When the F-1 development began, there was no rocket on the drawing board intended to use it, nor any mission defined which would require it. The Air Force had simply established that such an engine would be adequate to accomplish any military mission in the foreseeable future. When NASA took over responsibility for heavy launchers from the Air Force, the F-1 engine became central to the evolving heavy lifters envisioned for missions beyond Earth orbit. After Kennedy's decision to mount a manned lunar landing mission, NASA embarked on a furious effort to define how such a mission could be accomplished and what hardware would be required to perform it. The only alternative to heavy lift would be a large number of launches which assembled the Moon ship in Earth orbit, which was a daunting prospect at a time when not only were rockets famously unreliable and difficult to launch on time, but nobody had ever so much as attempted rendezvous in space, no less orbital assembly or refuelling operations.

With the eventual choice of lunar orbit rendezvous as the mission mode, it became apparent that it would be possible to perform the lunar landing mission with a single launch of a booster with 7.5 million pounds of sea level thrust, which could be obtained from a cluster of five F-1 engines (which by that time NASA had specified as 1.5 million pounds of thrust). From the moment the preliminary design of the Saturn V was defined until Apollo 11 landed on the Moon, the definition, design, testing, and manufacturing of the F-1 engine was squarely on the critical path of the Apollo project. If the F-1 did not work, or was insufficiently reliable to perform in a cluster of five and launch on time in tight lunar launch windows, or could not have been manufactured in the quantities required, there would be no lunar landing. If the schedule of the F-1 slipped, the Apollo project would slip day-for-day along with its prime mover.

This book recounts the history, rationale, design, development, testing, refinement, transition to serial production, integration into test articles and flight hardware, and service history of this magnificent machine. Sadly, at this remove, some of the key individuals involved in this project are no longer with us, but the author tracked down those who remain and discovered interviews done earlier by other researchers with the departed, and he stands back and lets them speak, in lengthy quotations, not just about the engineering and management challenges they faced and how they were resolved, but what it felt like to be there, then. You get the palpable sense from these accounts that despite the tension, schedule and budget pressure, long hours, and frustration as problem after problem had to be diagnosed and resolved, these people were having the time of their lives, and that they knew it at the time and cherish it even at a half century's remove. The author has collected more than a hundred contemporary photographs, many in colour, which complement the text.

A total of sixty-five F-1 engines powered 13 Saturn V flight vehicles. They performed with 100% reliability.

January 2012 Permalink

Zubrin, Robert. The Case for Space. Amherst, NY: Prometheus Books, 2019. ISBN 978-1-63388-534-9.
Fifty years ago, with the successful landing of Apollo 11 on the Moon, it appeared that the road to the expansion of human activity from its cradle on Earth into the immensely larger arena of the solar system was open. The infrastructure built for Project Apollo, including that in the original 1963 development plan for the Merritt Island area could support Saturn V launches every two weeks. Equipped with nuclear-powered upper stages (under active development by Project NERVA, and accommodated in plans for a Nuclear Assembly Building near the Vehicle Assembly Building), the launchers and support facilities were more than adequate to support construction of a large space station in Earth orbit, a permanently-occupied base on the Moon, exploration of near-Earth asteroids, and manned landings on Mars in the 1980s.

But this was not to be. Those envisioning this optimistic future fundamentally misunderstood the motivation for Project Apollo. It was not about, and never was about, opening the space frontier. Instead, it was a battle for prestige in the Cold War and, once won (indeed, well before the Moon landing), the budget necessary to support such an extravagant program (which threw away skyscraper-sized rockets with every launch), began to evaporate. NASA was ready to do the Buck Rogers stuff, but Washington wasn't about to come up with the bucks to pay for it. In 1965 and 1966, the NASA budget peaked at over 4% of all federal government spending. By calendar year 1969, when Apollo 11 landed on the Moon, it had already fallen to 2.31% of the federal budget, and with relatively small year to year variations, has settled at around one half of one percent of the federal budget in recent years. Apart from a small band of space enthusiasts, there is no public clamour for increasing NASA's budget (which is consistently over-estimated by the public as a much larger fraction of federal spending than it actually receives), and there is no prospect for a political consensus emerging to fund an increase.

Further, there is no evidence that dramatically increasing NASA's budget would actually accomplish anything toward the goal of expanding the human presence in space. While NASA has accomplished great things in its robotic exploration of the solar system and building space-based astronomical observatories, its human space flight operations have been sclerotic, risk-averse, loath to embrace new technologies, and seemingly more oriented toward spending vast sums of money in the districts and states of powerful representatives and senators than actually flying missions.

Fortunately, NASA is no longer the only game in town (if it can even be considered to still be in the human spaceflight game, having been unable to launch its own astronauts into space without buying seats from Russia since the retirement of the Space Shuttle in 2011). In 2009, the commission headed by Norman Augustine recommended cancellation of NASA's Constellation Program, which aimed at a crewed Moon landing in 2020, because they estimated that the heavy-lift booster it envisioned (although based largely on decades-old Space Shuttle technology) would take twelve years and US$36 billion to develop under NASA's business-as-usual policies; Constellation was cancelled in 2010 (although its heavy-lift booster, renamed. de-scoped, re-scoped, schedule-slipped, and cost-overrun, stumbles along, zombie-like, in the guise of the Space Launch System [SLS] which has, to date, consumed around US$14 billion in development costs without producing a single flight-ready rocket, and will probably cost between one and two billion dollars for each flight, every year or two—this farce will probably continue as long as Richard Shelby, the Alabama Senator who seems to believe NASA stands for “North Alabama Spending Agency”, remains in the World's Greatest Deliberative Body).

In February 2018, SpaceX launched its Falcon Heavy booster, which has a payload capacity to low Earth orbit comparable to the initial version of the SLS, and was developed with private funds in half the time at one thirtieth the cost (so far) of NASA's Big Rocket to Nowhere. Further, unlike the SLS, which on each flight will consign Space Shuttle Main Engines and Solid Rocket Boosters (which were designed to be reusable and re-flown many times on the Space Shuttle) to a watery grave in the Atlantic, three of the four components of the Falcon Heavy (excluding only its upper stage, with a single engine) are reusable and can be re-flown as many as ten times. Falcon Heavy customers will pay around US$90 million for a launch on the reusable version of the rocket, less than a tenth of what NASA estimates for an SLS flight, even after writing off its enormous development costs.

On the heels of SpaceX, Jeff Bezos's Blue Origin is developing its New Glenn orbital launcher, which will have comparable payload capacity and a fully reusable first stage. With competition on the horizon, SpaceX is developing the Super Heavy/Starship completely-reusable launcher with a payload of around 150 tonnes to low Earth orbit: more than any past or present rocket. A fully-reusable launcher with this capacity would also be capable of delivering cargo or passengers between any two points on Earth in less than an hour at a price to passengers no more than a first class ticket on a present-day subsonic airliner. The emergence of such a market could increase the demand for rocket flights from its current hundred or so per year to hundreds or thousands a day, like airline operations, with consequent price reductions due to economies of scale and moving all components of the transportation system down the technological learning curve.

Competition-driven decreases in launch cost, compounded by partially- or fully-reusable launchers, is already dramatically decreasing the cost of getting to space. A common metric of launch cost is the price to launch one kilogram into low Earth orbit. This remained stubbornly close to US$10,000/kg from the 1960s until the entry of SpaceX's Falcon 9 into the market in 2010. Purely by the more efficient design and operations of a profit-driven private firm as opposed to a cost-plus government contractor, the first version of the Falcon 9 cut launch costs to around US$6,000/kg. By reusing the first stage of the Falcon 9 (which costs around three times as much as the expendable second stage), this was cut by another factor of two, to US$3,000/kg. The much larger fully reusable Super Heavy/Starship is projected to reduce launch cost (if its entire payload capacity can be used on every flight, which probably isn't the way to bet) to the vicinity of US$250/kg, and if the craft can be flown frequently, say once a day, as somebody or other envisioned more than a quarter century ago, amortising fixed costs over a much larger number of launches could reduce cost per kilogram by another factor of ten, to something like US$25/kg.

Such cost reductions are an epochal change in the space business. Ever since the first Earth satellites, launch costs have dominated the industry and driven all other aspects of spacecraft design. If you're paying US$10,000 per kilogram to put your satellite in orbit, it makes sense to spend large sums of money not only on reducing its mass, but also making it extremely reliable, since launching a replacement would be so hideously expensive (and with flight rates so low, could result in a delay of a year or more before a launch opportunity became available). But with a hundred-fold or more reduction in launch cost and flights to orbit operating weekly or daily, satellites need no longer be built like precision watches, but rather industrial gear like that installed in telecom facilities on the ground. The entire cost structure is slashed across the board, and space becomes an arena accessible for a wide variety of commercial and industrial activities where its unique characteristics, such as access to free, uninterrupted solar power, high vacuum, and weightlessness are an advantage.

But if humanity is truly to expand beyond the Earth, launching satellites that go around and around the Earth providing services to those on its surface is just the start. People must begin to homestead in space: first hundreds, then thousands, and eventually millions and more living, working, building, raising families, with no more connection to the Earth than immigrants to the New World in the 1800s had to the old country in Europe or Asia. Where will they be living, and what will they be doing?

In order to think about the human future in the solar system, the first thing you need to do is recalibrate how you think about the Earth and its neighbours orbiting the Sun. Many people think of space as something like Antarctica: barren, difficult and expensive to reach, unforgiving, and while useful for some forms of scientific research, no place you'd want to set up industry or build communities where humans would spend their entire lives. But space is nothing like that. Ninety-nine percent or more of the matter and energy resources of the solar system—the raw material for human prosperity—are found not on the Earth, but rather elsewhere in the solar system, and they are free for the taking by whoever gets there first and figures out how to exploit them. Energy costs are a major input to most economic activity on the Earth, and wars are regularly fought over access to scarce energy resources on the home planet. But in space, at the distance Earth orbits the Sun, 1.36 kilowatts of free solar power are available for every square metre of collector you set up. And, unlike on the Earth's surface, that power is available 24 hours a day, every day of the year, and will continue to flow for billions of years into the future.

Settling space will require using the resources available in space, not just energy but material. Trying to make a space-based economy work by launching everything from Earth is futile and foredoomed. Regardless of how much you reduce launch costs (even with exotic technologies which may not even be possible given the properties of materials, such as space elevators or launch loops), the vast majority of the mass needed by a space-based civilisation will be dumb bulk materials, not high-tech products such as microchips. Water; hydrogen and oxygen for rocket fuel (which are easily made from water using electricity from solar power); aluminium, titanium, and steel for structural components; glass and silicon; rocks and minerals for agriculture and bulk mass for radiation shielding; these will account for the overwhelming majority of the mass of any settlement in space, whether in Earth orbit, on the Moon or Mars, asteroid mining camps, or habitats in orbit around the Sun. People and low-mass, high-value added material such as electronics, scientific instruments, and the like will launch from the Earth, but their destinations will be built in space from materials found there.

Why? As with most things in space, it comes down to delta-v (pronounced delta-vee), the change in velocity needed to get from one location to another. This, not distance, determines the cost of transportation in space. The Earth's mass creates a deep gravity well which requires around 9.8 km/sec of delta-v to get from the surface to low Earth orbit. It is providing this boost which makes launching payloads from the Earth so expensive. If you want to get to geostationary Earth orbit, where most communication satellites operate, you need another 3.8 km/sec, for a total of 13.6 km/sec launching from the Earth. By comparison, delivering a payload from the surface of the Moon to geostationary Earth orbit requires only 4 km/sec, which can be provided by a simple single-stage rocket. Delivering material from lunar orbit (placed there, for example, by a solar powered electromagnetic mass driver on the lunar surface) to geostationary orbit needs just 2.4 km/sec. Given that just about all of the materials from which geostationary satellites are built are available on the Moon (if you exploit free solar power to extract and refine them), it's clear a mature spacefaring economy will not be launching them from the Earth, and will create large numbers of jobs on the Moon, in lunar orbit, and in ferrying cargos among various destinations in Earth-Moon space.

The author surveys the resources available on the Moon, Mars, near-Earth and main belt asteroids, and, looking farther into the future, the outer solar system where, once humans have mastered controlled nuclear fusion, sufficient Helium-3 is available for the taking to power a solar system wide human civilisation of trillions of people for billions of years and, eventually, the interstellar ships they will use to expand out into the galaxy. Detailed plans are presented for near-term human missions to the Moon and Mars, both achievable within the decade of the 2020s, which will begin the process of surveying the resources available there and building the infrastructure for permanent settlement. These mission plans, unlike those of NASA, do not rely on paper rockets which have yet to fly, costly expendable boosters, or detours to “gateways” and other diversions which seem a prime example of (to paraphrase the author in chapter 14), “doing things in order to spend money as opposed to spending money in order to do things.”

This is an optimistic and hopeful view of the future, one in which the human adventure which began when our ancestors left Africa to explore and settle the far reaches of their home planet continues outward into its neighbourhood around the Sun and eventually to the stars. In contrast to the grim Malthusian vision of mountebanks selling nostrums like a “Green New Deal”, which would have humans huddled on an increasingly crowded planet, shivering in the cold and dark when the Sun and wind did not cooperate, docile and bowed to their enlightened betters who instruct them how to reduce their expectations and hopes for the future again and again as they wait for the asteroid impact to put an end to their misery, Zubrin sketches millions of diverse human (and eventually post-human, evolving in different directions) societies, exploring and filling niches on a grand scale that dwarfs that of the Earth, inventing, building, experimenting, stumbling, and then creating ever greater things just as humans have for millennia. This is a future not just worth dreaming of, but working to make a reality. We have the enormous privilege of living in the time when, with imagination, courage, the willingness to take risks and to discard the poisonous doctrines of those who preach “sustainability” but whose policies always end in resource wars and genocide, we can actually make it happen and see the first steps taken in our lifetimes.

Here is an interview with the author about the topics discussed in the book.

This is a one hour and forty-two minute interview (audio only) from “The Space Show” which goes into the book in detail.

June 2019 Permalink