Astronaut Training and Mission Simulation

Just before I arrived at NASA, in April 1963 the United States Geological Sur­vey had reached an agreement with the Manned Spacecraft Center to start a geological training program for the astronauts. Ellington Air Force Base, a few miles west of the proposed location for the main MSC campus and home of the NASA astronaut air force, was selected as the site for this rump USGS office. Gene Shoemaker chose Dale Jackson, a former marine, to lead this effort, thinking his background would allow him to mesh successfully with the astro­nauts, who were all military pilots. Until that time the astronauts were not perceived as enthusiastic about studying geology, in view of their other pressing duties. By the time I joined NASA, stories were already circulating that some MSC staff members and Jackson’s small team did not agree on who was to call the shots on this important function. MSC staffers believed they should be in charge, although USGS had been given this mandate by NASA headquarters. Adding to the problem, the newly hired MSC staffers assigned to work with Jackson’s people did not have as much experience as Jackson’s staff, yet he agreed to include them in the training. As in other areas I have described, MSC had a pronounced fear of being left out of important assignments related to Apollo science and tried whenever possible to monopolize these roles.

In spite of the friction between the two staffs, Jackson plowed ahead with his duties and devised classroom and fieldwork courses in basic geologic princi­ples, mineralogy, and petrology. With the astronaut office’s approval, the syl­labus called for fifty-eight hours of classroom lectures and four field trips. The fifty-eight hours of ‘‘geology’’ training were part of an overall classroom syl­labus of 239 hours designed to prepare the astronauts for the upcoming Gemini flights.1 The geology training was not related to the upcoming Gemini flights, the astronauts’ primary concern at that time, and would not have real value unless they were selected as Apollo crewmen. Thus it was not universally em­braced, especially by some of the original seven and the second and third astronaut classes. Eventually, however, it became accepted as an essential box to be checked off if one hoped to be chosen for a Moon mission. It was anticipated that after crews were selected for the lunar landing missions, five additional series of follow-on lectures and field trips would be scheduled.

By 1967, one hundred hours of classroom lectures and ten field trips became the requirement for astronaut geology training. This training, and then the mission simulations, would become more and more rigorous and realistic as the program matured and simulations were scheduled using prototype and final design equipment and tools.

Three weeks after joining NASA in September 1963, I attended my first demonstration of a prototype Apollo space suit at MSC. The demonstration and briefing were done under the auspices of MSC’s Crew Systems Division. Hamilton Standard had been awarded the overall contract to develop the Apollo space suit and backpack, with International Latex, its subcontractor, responsible for the suit itself. This was my first opportunity to see the current state of the art in space suits. The prototype Apollo suit we were to see demon­strated was the latest amalgamation of this technology, plus modifications added by the Crew Systems staff, which had the ability (or expertise) to second – guess the contractor and make its own adaptations when appropriate. At this point two types of suits were under consideration: a ‘‘soft suit’’ made of multi­ple layers of nylon and other material and a ‘‘hard suit’’ to be made of some type of hard plastic or honeycombed aluminum material. This was a ‘‘soft suit’’ demonstration, the preferred approach.

A test engineer wearing the suit went through a series of mobility exercises for the assembled throng. Some movements he could carry out easily; others were more difficult or almost impossible. Bob Fudali and Noel Hinners of Bellcomm also attended the demonstration and filed a detailed report on what they had observed. They wrote: ‘‘All in all, it looks as if mobility will be rather low (even in improved suits) and that the astronauts will not travel far from the LEM without additional mechanical aids. [Their] ability to set up equipment and perform experiments on the surface will also be quite limited unless strik­ing changes are made in future suits.’’2 I also reported in a memo to my office what I had seen and what I believed were the deficiencies in the design.

My first exposure to astronaut training and simulation came at the end of August 1964 with a trip to Bend, Oregon. At this early date many had ques­tioned the astronauts’ ability to carry out meaningful scientific observations and work on the lunar surface while encumbered by the available space suits. I was one of the skeptics, based on the earlier space suit demonstration at MSC. My report on the 1963 demonstration had gotten back to Max Faget’s office at MSC and was considered so negative that when MSC found out I would be attending the Bend simulation, Faget sent a telegram to Tom Evans disinviting me. Ed Andrews told me to ignore the telegram and go anyway.

The Bend simulation, supported by several MSC offices, was designed around a space-suited astronaut, Walt Cunningham, alternating with two MSC technicians in space suits. They would work at several locations, using a few rudimentary field tools, and at the same time report what they were doing and seeing. The Bend location was chosen because it seemed like a good terrestrial analogue of what the astronauts would find on the Moon. It consisted of three types of volcanic terrain. One site was primarily a field of basaltic extrusives, jagged and rough and in places containing pieces of obsidian. MSC, it was rumored, was considering using the area as a permanent simulation site. Gover­nor Mark O. Hatfield (not yet a senator) and the press had been invited to witness parts of the simulation, and the exercise rapidly turned into a major public relations gaffe.

During the simulations, Walt wore the prototype Apollo space suit demon­strated less than a year earlier, with a few improvements including a new back­pack. It was the best suit available at the time. Together the suit and backpack and a bulky white overgarment weighed more than a hundred pounds. It was a blazing hot day, uncomfortable even for those of us just standing and watching in shirtsleeves. Walt’s suit was fitted out with a new water-cooled inner garment, best described as a pair of long johns with a network of thin plastic tubes sewn on. Cold water circulating through the tubes was supposed to keep him from overheating. It didn’t. His visor often fogged over, and he had trouble seeing where he was going.

One slope he tried to climb was covered with pieces of razor-sharp obsidian, and as might be expected, he tripped and sliced a hole in one of his gloves. Before this he had tried to use a geologic hammer and scoop to pick up samples. Both tasks were awkward in such a garment, but to make matters worse he had to carry the tools in one hand or hung at his waist and at the same time manipulate either a “walker” or a ‘‘Jacobs staff’’ that was supposed to help him conquer this rough terrain. At every stop he would put down the walker or staff and begin his next task. No matter how hard he tried, every action looked difficult. Whenever he bent over he tended to lose his balance because the suit was not designed to bend easily at the waist, a deficiency we had noted a year earlier. After he fell and cut his glove he continued to tumble down the slope and was saved from injury only by two technicians standing nearby just in case. All in all, it was a simulation disaster, which the local press reported the next day in large headlines.

By the end of the simulation, with a short rest after his fall while the tear in his glove was repaired (‘‘duck tape’’ helped get us to the Moon), Walt attributed his problems to his fogged-over visor and other suit limitations. He described the scene to his superiors back in Houston as a ‘‘Roman holiday,’’ referring to the swarming photographers eagerly taking pictures of his pratfall. Bob Fudali of Bellcomm also was there to observe the simulation. In his report he noted that ‘‘predicting the mobility of an astronaut on the lunar surface from these tests would be a serious error.’’3 My report to my office also retold Cunning­ham’s mishaps, and when copies of our memos were brought to his attention, he came to associate us with his bad press. The main points of our memos had been to argue for a suit that would make the astronauts more mobile and for better-designed tools, not to criticize Walt’s efforts. This simulation was an important factor that led him to caution us at the Falmouth summer con­ference not to overload the astronauts with lunar surface science tasks. Later I was able to explain my position to him and we became good working partners, though Walt never quite forgot his embarrassing Oregon experience.

My report also addressed the disadvantage of having such a large public attendance at simulations where many new things would be tried for the first time. I recommended that future simulations be done at Flagstaff, where we were beginning to set up good facilities and where attendance might be con­trolled. I had, of course, an additional motivation: to legitimize the role USGS was playing in our post-Apollo simulations and put the staff in a position to more strongly influence what would be done for Apollo. Will Foster and E. Z. Gray agreed with my suggestion, and each sent a memo to George Mueller recommending that Flagstaff be the future site for simulations.4 The Office of Space Medicine also sided with our observations and recommended policies to guide future simulations, including that astronauts ‘‘not be used as test sub­jects’’ unless they would make some unique contribution.5 Mueller forwarded these memos to MSC. He got back a letter from George Low, deputy director at MSC, disagreeing with Foster and Gray on their recommendation to conduct future field simulations requiring special terrain at Flagstaff and claiming there was no intent to set up a ‘‘lunar training camp’’ at Bend.6 This last statement played down Governor Hatfield’s comments while he was at the simulation that he supported having such a ‘‘camp’’ at Bend. This seemed to confirm the rumors we had heard that MSC had indeed made some preliminary overtures. It was clear that Low was telling Mueller they intended to do their own thing, especially when dealing with USGS.

Low’s response prompted Foster to send Mueller another memo to clear the air; he said that his earlier memo was not intended as a criticism of MSC but repeated his concern that pressure was being exerted on NASA to establish a training facility at Bend.7 To put an end to this internal bickering, Mueller wrote to Bob Gilruth, the MSC center director, ‘‘It is my desire that the Centers work closely with the USGS. . . and that there be no unnecessary duplication of field simulation activities,” and he sent an identical letter to Wernher von Braun at Marshall Space Flight Center.8 This exchange, unfortunately, only deepened the growing animosity between MSC and our headquarters-USGS team.

As field geology training picked up speed and our post-Apollo studies pro­gressed, we were constantly trying to find sites that would demonstrate terrains similar to those we expected the astronauts to encounter on the Moon. USGS already had a selection of sites it used at different stages in the training program, depending on the objective. Training trips took the astronauts to many distant places, both in the United States and overseas. But as our understanding of the Moon grew from pictures returned by Ranger, Surveyor, and Lunar Orbiter, new sites that could mimic the lunar surface were in demand for both Apollo and post-Apollo mission planning.

In May 1964 Bill Henderson, Don Elston, William Fischer of USGS, and I went hunting for sites that might be suitable for simulating longer missions and lunar base activities. Final reports from Bill Henderson’s Lunar Exploration Systems for Apollo (LESA) lunar base studies were due in nine months. Interim reports were already suggesting a broad range of undertakings that could be carried out at a base, and we used these early reports as a starting point for planning lunar base simulations. In those heady days we were thinking big; a lunar base program would undoubtedly be announced in the near future, to follow the successful Apollo missions. Until this time simulations for post – Apollo missions had been conducted exclusively near Flagstaff. We were look­ing for one or more large sites, not too remote and preferably on government property, where we could expect to find support for the lunar base simulations, which we anticipated would be complex. We drew up a list of potential loca­tions, obtained photographs and other background material, and reduced the large number of candidates to a short list.

We went first to the Atomic Energy Commission’s Nevada Test Site (NTS), where a series of surface and subsurface atomic and high energy chemical explosive tests had pockmarked the landscape with craters of all sizes. The local AEC manager was interested in our proposal, and though the site had restricted access, some sections could be made available for training. We were given a helicopter overflight, and from the air there was no question that it appeared moonlike. One crater, called Sedan, was especially impressive. Formed by a 104 kiloton explosive, the crater was 320 feet deep and 1,280 feet across. Flying over it at low altitude reminded me of standing on the rim of Meteor Crater in Arizona, for it had many of the same characteristics. After we landed we toured the site by truck to get a closer view. When we got out of the truck at the first stop, we discovered a major problem; we had to put on white coveralls and boots because the surface soil was still slightly radioactive; the atomic clocks of some of the products of the nuclear explosions were still ticking. We should have expected this situation, but when we made our calls to set up the tour, the fact was not mentioned. We looked at each other and rolled our eyes, then after a few short excursions we thanked our hosts politely and left.

Our second stop was China Lake, a large navy test range in southern Califor­nia. We studied a large-scale map of the range at the headquarters building and selected a few spots for a close-up truck survey. The range was vast (1.1 million acres), with lots of room for the many exercises we were hoping to conduct. Although it was not as Moonlike as NTS, vegetation was sparse and there were many interesting geological formations that could simulate lunar conditions. We toured the range by truck and agreed that it looked like a good site, and the commanding officer seemed willing to accommodate us. The test range also included many shops, hangars, and other facilities that we would need to support long-staytime simulations. They could be made available, we were told, with appropriate compensation.

From China Lake we next visited Fort Huachuca, Arizona. After a meeting with the commanding general, who assured us of his interest, the army also provided a helicopter overflight, followed by a series of briefings on facilities and other advantages of working there. They were definitely selling: perhaps they saw reduced budgets in their future and thought this new use might offset these reductions. This army proving ground was beyond question isolated. The Huachuca Mountains formed the western border of the fort, and a variety of volcanic terrains could be found within its boundaries. Although the region was semiarid, it was a ‘‘green desert.’’ Most of the ground was covered with cactus, including cholla, palo verde, and other types of plant life common to the area; it was beautiful, but we thought it would be too difficult to cope with continuously for sustained long-distance walking and vehicular simulations.

Our final stop was the White Sands Missile Range in south-central New Mexico. It was similar in many respects to China Lake. There was lots of space, some areas had Moonlike terrain, and there were good support facilities. NASA was already using some of the range, so we would not be unwelcome guests. It was perhaps the best of the sites we visited. As events unfolded, we never had to make a choice. Lunar base funding and planning came to an end about a year later, and our more modest post-Apollo simulations were all carried out near Flagstaff.

We continued to look for additional Apollo training sites, however, and a new tool became available to assist us. On each Gemini flight the astronauts took photographs of the Earth’s surface with handheld Hasselblad cameras. Many showed areas never before well documented with aerial photographs. For each flight Paul Lowman, with his coinvestigator Herbert Tiedemann at MSC, had designated points of special interest that the crew should try to photo­graph, time permitting. Gemini missions were launched due east from Kennedy Space Center to take full advantage of the extra boost from the Earth’s rotation; thus their flight paths repeatedly covered all of the Earth’s surface from 28.5° north latitude to 28.5° south. One of the benefits of repeating the launch inclination was that it was possible to rephotograph the designated areas when the photos from earlier missions were of poor quality or were not taken. This also allowed some stereoscopic coverage where the photos overlapped.

Using these photos, Paul and I searched for other potential training sites. Each Gemini photo typically covered an area of some 3,500 square miles, with the oblique photos covering even more—an unprecedented continuous view of the Earth’s surface. In the typical aerial survey, an average frame might cover less than ten square miles. Conventional photographic coverage of the large areas included in a typical Gemini frame would require constructing photo mosaics, with trained photogrammetrists piecing together many separate pho­tographs. Having used such products in our geological pasts, we knew that no matter how skillfully fabricated, photo mosaics always introduced false infor­mation in the finished maps. A geologist could be misled by something that looked like a stream or valley or some geological feature such as a fault but was really an edge between two photos.

Features never fully photographed before the Gemini missions, such as the Richat structure in Mauritania, that might be the result of large meteorite impacts were of special interest because they might provide not only training sites but also the opportunity to learn more about impact processes. In 1965 only a few well-documented impact craters were known throughout the world, and many of them were so obscured by erosion that they were not well suited as training sites. Thus we were constantly trying to find more examples that we could study or use to train the astronauts.

A few of the Gemini photos had been published in National Geographic, Life, newspapers, and other publications, but the vast majority had not been seen by the general public. In his spare time Paul had been carefully cataloging the pictures and interpreting their geologic content. It occurred to us that these new views of the Earth might interest companies exploring remote parts of the world. So far, no commercial interest had been shown. If we could get a positive response, it would support NASA’s proposed Earth orbital remote sensing program-just in an early planning stage—and perhaps persuade NASA man­agement to accelerate this program.

In May 1966 I called Mobil Oil in New York and talked to my old boss, James Roberts, who had been transferred after I left Colombia, first to Venezuela and then to Mobil headquarters. I explained what we had and what we thought would be the potential benefits and applications of space photography. He said he was interested in seeing the photographs and agreed to set up a meeting with some of the Mobil Exploration staff, the unit responsible for finding new oil fields. A few weeks later Paul and I flew to New York to show the Gemini photos to their first commercial audience. We brought to the briefing some of the best examples of geological features photographed by the astronauts; mountain ranges in the southern Sahara (Mobil was heavily involved in exploring remote areas in Libya and Tunisia) and clear pictures of structures in Iran of the type petroleum geologists looked for (anticlines and synclines). I knew Mobil had several field parties working in Iran at that time, because before I left Colombia Iran was a possible new destination for me. We also included a few spectacular views of the Andes and the Himalayas. We felt sure there were no aerial photo­graphs of some of these areas, and this would be the first time Mobil had such views available. We thought they would be impressed.

We were wrong. For whatever reasons, the staff members Roberts brought to our meeting showed little interest. They said they had, or could get, enough conventional coverage so that space photographs were not needed. This re­sponse mystified us. Perhaps they thought an endorsement would leave them open to providing financial support for an undertaking with an uncertain future. We will never know what might have happened if Mobil had been enthusiastic. Like other programs that were struggling to get started at this time, the Earth orbital observation program limped along, in part because there was no strong commercial interest. It would be many years before the unmanned Landsat program and Skylab would be launched.

Our search for terrestrial impact structures took us on two trips, one back to Colombia in April 1964 and another to Peru in June 1968. We visited Colombia to study a small circular structure of unknown origin, Lake Guatavita, high in the eastern cordillera of the Andes, some thirty miles north of Bogota. Lake Guatavita was an intriguing and well-known feature; at the time of the Spanish conquest it was rumored that the Chibcha Indians, who lived on the high plateau that surrounds what is now Bogota, used the lake for special cere­monies. It was said that the local chief would cover himself in gold dust every year and then bathe in the water, accompanied by other sacrificial ceremonies. The Spanish had dredged the lake and attempted to drain it in hopes of finding sunken treasure. A modern attempt, again unsuccessful, had also been made to drain the lake after several marvelously intricate gold artifacts were recovered from the bottom. Geological study had failed to come up with a satisfactory explanation of the lake’s almost perfectly circular shape; one suggestion was that it was created by an impact, but no proof had been reported. I had visited the lake while living in Colombia and was aware of its history and the impact theory.

Now that there was better understanding of how to identify an impact crater in the field, Paul and I developed a field study plan for making a quick assess­ment of the lake and submitted it for approval. The estimated cost of the trip for the two of us, including all expenses, was $1,000. In the memos that went back and forth before approval was given, a number of interesting comments were appended to the routing slips. The most humorous was one made by George Mueller’s special assistant, Paul Cotton: ‘‘George, this is the slickest justification for a boondoggle I have ever seen. As long as we have this kind of resourcefulness, we should be confident of reaching the moon and planets.’’ A second staff comment to Mueller was that approval should be given only if we included an astronaut. We were in favor of this recommendation, but it was soon shot down as taking too much valuable astronaut time. Our ‘‘resourceful­ness’’ was rewarded, and the trip was approved.

Our plan was to quickly survey the lake’s immediate surroundings looking for evidence of impact in the form of shatter cones or other impact debris such as ejecta, glass, or meteorite material. For two days we tramped around the half­mile-diameter lake picking up samples, taking pictures, and making a few measurements. We could find no evidence of an impact. This left us in a quandary: How should we report our results when there was so little to report? We felt sure that thin-section study of our samples would only confirm our field observations that the lake was not the result of impact. We went back to my old Mobil office in Bogota to examine more closely what was known, geologically, of the immediate area. Based on the published literature, we concluded that since we could find no evidence of an impact the lake was probably formed when the surface rock collapsed over a small salt dome that had been dissolved by groundwater. Thick salt deposits were known to exist in the underlying formations, and a complete cathedral had been carved below ground from the salt at Zipaquira, a short distance away. And so we reported our findings.9

When E. Z. Gray forwarded our report to Mueller we received a short handwritten acknowledgment: ‘‘I doubt if the returns were worth the time and money. Do you agree?’’ Gray wrote back: ‘‘What value do you place on develop­ing an organization? I am a firm believer in learning by doing. I think this trip was worthwhile.” Although it was only a small incident in a rapidly accelerating major national undertaking, this story provides a measure of the attention to detail demonstrated by senior management and at the same time the freedom of action they allowed their staffs. Such management competence, and such security in their abilities, may have had no equal in a government program before or after and was, I believe, instrumental in Apollo’s success.

The Peru trip was instigated by our study of the photographs returned by

Gemini 9. During the flight the astronauts had photographed the Andes from Chile to Colombia. At the point where the mountain chain turns from a mostly north-south direction to the northwest near Lake Titicaca in southern Peru, we observed several large circular structures, each having a diameter of thirty miles or more. Were they created by impacts or by some other mechanism?

After Paul and I found the circular structures on the Gemini photographs, we tried to determine if they had been discussed in the geological literature. We found no citations. Such large structures, if formed by impacts, would be a major discovery. We could see many large impact craters on the Moon, and by this time we had in hand the detailed Lunar Orbiter photographs that showed some of the fine structure associated with large impacts. We knew of no impact craters of this size on Earth, although we were sure that, like those on the Moon, they had been made during the planet’s early history. The Ries Kessel structure in Germany, about fifteen miles in diameter, which was used as an astronaut training site, was the largest confirmed terrestrial impact feature known at that time. The Vredefort Dome in South Africa, some twenty-five miles across, was potentially a larger example but was yet to be studied in detail. Many aspects of the large lunar craters were intriguing, especially their central peaks. Only large lunar craters had such peaks. Why did they exist? Did they reflect the thickness of the lunar crust or some other unknown phenomenon? The Gemini photos showed that the large circular structures in Peru had mountains in their cen­ters. We started to lay plans to visit Peru and try to answer our questions on the origin of these features.

As our planning progressed, Paul could see it would be difficult for him to make the trip; he had returned to Goddard Space Flight Center and new duties. I continued to pursue the idea and finally received permission to go from my new boss, Lee Scherer. In preparation I had been in contact with the United States and Peruvian embassies as well as the Peruvian Geological Survey and was assured of their cooperation. From the Defense Intelligence Agency I had obtained aerial photographs of the area taken in 1955 so I could plot our findings in the field. Interestingly, these relatively high resolution individual photographs gave no indication of the structures, and a photomosaic made from these photos also failed to show them. The advantage of the small-scale space photos, which covered a large area without distortion, was clear. In addition to these rather formal arrangements, I received an unexpected bonus. A NASA colleague, Rollin Gillespie, who worked in the Planetary Missions

Office, was interested in joining me. His son Alan, who was majoring in geology at Stanford, was also interested; so Rollin, at his own expense, offered to meet me in Lima and accompany me along with several Stanford students.

I arrived in Lima on June 15 sans baggage and field equipment, lost some­where en route. Rollin and his group had arrived several days before and had been in touch with the Peruvian Geological Survey. He had already made arrangements for two Land Rovers and for drivers, guides, translators (Spanish to Quechua), and three Peruvian geologists to accompany us. This saved us several days, since I arrived on the weekend and could not have made such connections for two days. While waiting for my baggage we met with the minerals attache at the United States embassy and with several other organiza­tions that were conducting mining operations in the area, and they supplied important information about the conditions we would encounter. An engineer at the Madrigal Mining Company told us they were working several large copper and silver mines in the center and on the flanks of two of the structures. This was encouraging; perhaps these circular features were similar to the Sud­bury structure in Canada, thought by some to be the remains of an impact crater, which was being mined for nickel, copper, and other metals.

Our plan was that Rollin and I would fly to Cuzco, where we would be joined two days later by the rest of the party and the Land Rovers and then travel south to the site. We flew to Cuzco on schedule and met, as we had arranged, with geologists at the National University of San Antonio to explain our project. They had never seen the Gemini photos and were excited by them. They were familiar with the region but had never realized these circular structures existed. While visiting at the university we received our first bad news. The rest of the party had been delayed in leaving Lima and would not arrive for several days. We decided to have them bypass Cuzco and meet us at Sicuani, a town near the base of the mountains. Before leaving the university I promised to stop on my way back to Lima and lecture to faculty and students on the Apollo program.

The next day Rollin and I took a bus to Sicuani, the only ‘‘gringos’’ on a bus filled to capacity with local passengers and all their baggage, some of it alive. It was essentially a straight shot through the Vilcanota Valley, which connects Cuzco to the altiplano that surrounds Lake Titicaca. Sicuani lay some eighty – five miles south of Cuzco by way of unpaved roads but with some spectacular scenery along the way. We arrived in Sicuani late in the afternoon and checked into the only hotel (warm water available every morning from 7:00 to 7:30). It was very cold. Sicuani is at an elevation of 12,000 feet, and there was no heat in the rooms, where we spent an uncomfortable night. By chance, while walking in the main plaza that first night, we met an American Carmelite priest who invited us to the parish house, where we discussed our plans with the assembled fathers. We then received our second round of bad news. They had visited the general area and told us it was not possible to drive in—it was too rough and there were no roads. We would have to rent horses. This would certainly slow up our exploration and add more time than I had available. They suggested we enlist the bishop’s support.

We met Bishop Hayes the next morning, and he was very helpful. Not only did he understand local politics and know who could ease the way, but he had a large, comfortable house (hot water all day) where he invited us to stay. We immediately agreed. The rest of our party arrived the next day, and we com­pleted our arrangements for renting horses and obtaining other equipment. With the delays in getting started my time in Peru was running out. I would be unable to travel to the structures and would have to depend on Rollin and the Stanford students, along with the Peruvian geologists, to complete the survey.

Returning to Cuzco by train, I stopped for the afternoon to deliver a lecture at the university. From Cuzco I flew back to Lima and then home. Back at NASA, I received a package from Professor Carlos Kalafatovich V. on the staff at the university in Cuzco. It contained several Peruvian newspaper clippings noting that scientists from NASA had visited the region and were interested in the mountains near Sicuani. According to the papers, which featured big black headlines that translated to ‘‘Flying Saucers Land in Canchis’’ (a small town near Sicuani), some of the local people interviewed were intimately familiar with those mountains. It seems that the locals knew of frequent visits by flying saucers that came to extract precious gems from somewhere in the mountains and take them back to their home planet. Now we knew what had attracted us to these structures.

On a more serious note, the party I left behind was not very successful. It was almost impossible to travel in the mountains, even using horses. They collected a few samples and took them back to Stanford for analysis. They found nothing unusual, and no sign of impact was observed in the mineralogy of the returned samples. The origin of the circular structures was not solved, and as far as I know the question is still open.

Backing up a bit, in September 1965 I participated in one of the astronaut training trips to Medicine Lake, California, a site near several small, complex volcanic features. By this time astronaut training trips were well organized by USGS and included prominent geologists who could lecture and teach the astronauts about the importance and subtleties of the locations selected and about their potential similarities to lunar features. This was the second two-day trip astronauts made to the area, and those on this particular trip were Russell ‘‘Rusty’’ Schweickart and Roger Chaffee. Roger was soon to be named to the crew selected to fly Apollo 1, scheduled to be the first manned flight of a Saturn rocket. Gene Cernan was also scheduled for this trip, but because of a hurricane threat he was delayed in Houston and unable to attend.

Roger Chaffee had come to the astronaut corps from the navy and held the rank of lieutenant commander. Since we were both jet pilots with many similar interests and experiences and had flown off some of the same class aircraft carriers, we hit it off immediately, and he became my truck mate for the training trip. I drove, and between scheduled stops and lectures I would fill him in on geological lore I thought he should know. But as I remember, we mostly swapped sea stories about night carrier landings and the idiosyncrasies of the planes we flew. He seemed to welcome the change of pace from his ‘‘normal’’ astronaut assignments, even though each day he was subjected to nonstop lectures and fieldwork while being force-fed textbook geology.

The team assembled for this trip consisted of ten people. Aaron Waters led the team and was to deliver the lectures and coordinate the trip itinerary. He was supported by nine helpers, including three USGS camp hands, two USGS geologists, two MSC geologists, and two MSC photographers. The astronauts’ doings were always well documented by photographs. Dick Allenby and I were also invited for this trip, so there were fourteen of us. We slept in one – or two – man tents and were up at dawn to complete each day’s tightly scheduled busi­ness. Breakfast was served around a campfire because the early morning hours were already chilly. At noon we had box lunches, and dinner was back at the campsite. This trip turned out to be especially memorable because William Rust, one of the USGS ‘‘camp hands’’ but in reality a technician, was the designated cook and an inveterate fisherman. Each morning, before any of us were awake, Bill would go to the lake and catch trout, then cook them for breakfast—a treat in any circumstances but for these few days a Washington bureaucrat’s delight.

Roger Chaffee’s attendance was especially significant and attested to the astronauts’ growing awareness of the importance of these trips as well as to Roger’s personal interest. Usually astronauts who would soon receive flight assignments could not take time off to attend to business other than that directly related to their flights, and there definitely was no geology to be done on Apollo 1. Roger enjoyed the training and was becoming an able field geolo­gist. I’m sure he hoped word of his new skills would get back to Deke Slayton and Al Shepard and put him in line for future Moon missions.

I told him I intended to submit my application for the next scientist – astronaut selection and hoped I would soon join him in the astronaut corps. Neither Roger’s flight nor my selection came to pass; less than two years later Roger died tragically in the Apollo 1 fire along with his two crewmates Virgil ‘‘Gus’’ Grissom and Edward White. Their deaths led directly to a major re­evaluation of how NASA was preparing for the Apollo missions, however, and the changes in the way NASA would do business ultimately ensured the pro­gram’s success.

Here is as good a place as any to relate my own experience in attempting to become an astronaut and give some idea of how scientist-astronauts were se­lected. Although I had been a military pilot, as were almost all the astronauts, I didn’t have a lot of jet hours; most of my flight time had been logged on propeller aircraft many years earlier. After working with the astronauts for a year and knowing their flight backgrounds, I could see that it would be virtually impossible for me to qualify in a typical selection process because I lacked current piloting experience. Then I heard that scientist-astronauts might be recruited. In April 1964 NASA asked the National Academy of Sciences to develop procedures for selecting them. Gene Shoemaker had lobbied for such a selection, and before he was diagnosed with Addison’s disease he had been considered a probable top choice when NASA finally got around to agreeing it needed such positions. Even after knowing he would not be selected, Gene continued to lobby, and his efforts, along with those of others in the science community, eventually paid off. I bided my time feeling that my best chance to qualify for the astronaut corps would be through the scientist-astronaut program.

When the call for applications was finally announced in October 1964, I quickly obtained the packet with the paperwork to be completed. It listed standards for such qualifications as age, height, and educational background.

Height! Maximum allowed height was six feet. I was six feet one. The age limit excluded anyone born before August 1, 1930. I was nine months overage. I made a few calls to see if these requirements were inflexible and found that they were. The height restriction was based on the dimensions of the Gemini cap­sules and the Apollo equipment then under design, which would not comfort­ably accommodate anyone over six feet. Greatly disappointed, I wrote to the National Academy of Sciences, the initial screening hurdle, to tell them I was interested but was disqualified because of my age and height, and that I hoped these restrictions might one day be changed so that I and others in my predica­ment could apply.

The good news about this first scientist-astronaut selection was that Jack Schmitt, then working on projects we were sponsoring at Flagstaff, made it all the way through, and he and five others became the first of this special group. Suddenly we were to have a strong advocate in Houston, someone who saw eye to eye with our concerns; but we would have to wait a year for his help while he trained to be a pilot.

I had written to the Academy with deliberate forethought. I felt sure there would be other scientist-astronaut selections. Our post-Apollo planning at that time called for extensive scientific experiments on the lunar surface, and quali­fied scientists would have to perform them to satisfy the scientific community. George Mueller had testified before Congress on these plans, and I knew he supported the need for additional scientist-astronauts. My letter, I hoped, would be retrieved at the next selection, showing my long-term interest in the program and perhaps influencing the selection criteria.

To give myself a better chance in the next selection, whenever it might be, I decided to apply for a pilot slot in one of the Navy Ready Reserve squadrons at nearby Andrews Air Force Base. My last flying experience had been with a navy reserve squadron in Denver while attending graduate school. No pilot openings were available at Andrews in 1964, so I joined an intelligence unit drilling once a month to get back in the Ready Reserve flow and learn through the grapevine where pilot assignments might be found.

This contact soon turned up a vacancy at Lakehurst Naval Air Station, and I quickly transferred to VS-751, an antisubmarine squadron, to resume flying after a seven-year layoff. A year and a half later, with new flying time under my belt, I persuaded a fighter squadron commander at Andrews who needed pilots to have me transferred, and I began the transition to the F8U Crusader. But the navy got wind of this behind the scenes activity; needing antisubmarine – qualified pilots, it rescinded my transfer and assigned me to VS-661 at Andrews. Although I was disappointed (I was looking forward to flying the Crusader, one of the navy’s best-ever fighters), the transfer had one redeeming factor. I would now fly out of Andrews and save the long monthly commute to Lakehurst. And at least I was flying and could hope that this would be a plus in the next selection.

In September 1966 the National Academy of Sciences announced the second scientist-astronaut selection. Accompanying the press release was a short state­ment by Gene Shoemaker, who would be chairman of the Academy’s selection panel: “Scientific investigations from manned space platforms and direct obser­vations on the Moon will initiate a new phase in man’s quest for knowledge. While such missions call for daring and courage of a rare kind, for the scientist they will also represent a unique adventure of the mind, requiring maturity and judgment of a high order.’’ Who could resist such a challenge? I thought that, with Gene as chairman and knowing several other members of his panel, I would have a real chance. It was rumored that this would be a larger class than the previous group of six, thus improving my odds. The Academy had been somewhat disappointed by the number of applications received for the first selection, although the six chosen had excellent qualifications, and thus the selection criteria were a little more relaxed the second time. The age and height limitations had not been changed, but this time the press release stated that “exceptions to any of the. . . requirements will be allowed in outstanding cases.’’ Perhaps now I had a chance. Could I qualify as an “outstanding case’’?

My application must have been one of the first received. As I remember, almost five thousand applications were screened for this second selection. Evi­dently there had been enough good publicity about the Apollo program in the interim to encourage many young scientists to want to be a part of it. About two hundred were selected for the next phase of physical and psychological exam­inations; I made the cut. We were divided into small groups and sent to the Air Force School of Aerospace Medicine at Brooks Air Force Base in San Antonio, where all astronaut candidates were screened.

We endured a week of prodding, blood work, and spinning, IQ, and many other tests, some of which were vividly shown in the movie The Right Stuff, though not with the same comic detail. (For a more complete account of what we experienced, read Mike Collins’s book Carrying The Fire.) While I was tilted upside down with my stomach filled with a barium solution, they discovered that I had a slight hiatal hernia; the muscles in my esophagus couldn’t hold all of the solution in my stomach. Because it was apparently a minor ailment and because, I assume, the other test results were good, I was sent to the Walter Reed Medical Center in Washington, D. C., for a second opinion. The examination at Walter Reed went well, and the examining doctor wrote a letter to NASA saying he did not consider the diagnosis disqualifying—that at the worst I might have to take an antacid to relieve any discomfort I might feel in zero gravity.

Where did this leave me? I couldn’t be sure, but I did have enough experi­ence to know that astronaut selections were secretive. I knew Deke Slayton and Al Shepard were involved, but I didn’t know who else. By this time I was acquainted with all the astronauts, including Al and Deke, but I wasn’t sure whether this was good or bad. I had been on field trips with them, from time to time I was invited to brief the astronauts on the plans for post-Apollo missions, and I was often in the astronaut office building to visit Jack Schmitt and other astronauts as well as the Crew Systems staff. I felt I had a good relationship with them, but perhaps my differences with some MSC managers might hamper my selection. In June I received the call I had been hoping for. I had made the final cut and was invited to Houston for the last interviews before a selection was made.

In June 1967 twenty-one candidates made this final visit. A few of them I knew from my week in San Antonio. Their backgrounds included almost all scientific disciplines, but as I read the list I saw I was the lone geologist, along with one geophysicist. Only two earth scientists! Most of the post-Apollo sci­ence activities we were planning had some earth science connection; I thought my selection was in the bag. The first scheduled activity after checking in was a ride in a T-38, the astronauts’ aircraft of choice, based at Ellington Air Force Base. This was a piece of cake. I flew the plane from the front seat with a NASA pilot (perhaps evaluator?) in the back seat. I did some simple maneuvers and a few snap rolls and generally showed off my flying skills. From what I read in the brief bios of the other candidates, I believed I was the only one with experience as a jet pilot. If this was a test, I must have passed. Next we took a ride in the MSC centrifuge; as I remember, they spun us up to about six gs while we performed a few simple exercises of hitting some light switches. Not a problem, and I suspect some of our future bosses were looking on through closed-circuit television to see how we did on the nearest thing to a stressful test.

After a few other briefings came the interview. I recall only four people in the room: Al, Deke, Bill Hess, and Charles Berry, who was head of the medical sciences office—‘‘the astronauts’ doctor.’’ All the questions were rather innocu­ous. Berry asked about the hiatal hernia, and since I had seen the Walter Reed report I told him that I hadn’t even known I had it until the test and that I didn’t think it would cause any trouble. The only question that stands out in my mind was the one Deke asked: ‘‘Don’t you think you’re too old to be an astronaut?’’ I was thirty-seven at the time and not the oldest of the final twenty-one candi­dates, but I knew I was over the advertised age allowance, so I had done a little homework. I answered, ‘‘I don’t think so; after all, I’m younger than Wally Schirra, and he’s still flying.’’ This brought a big laugh from all four inquisitors. Considering that Walter Schirra, then forty-three, was the only astronaut from the original seven to fly in all three programs—Mercury, Gemini, and Apollo— my answer was evidently on the mark. That ended the interview, and Al said he would give me a call. I thought my selection was now only a formality. That afternoon I did some preliminary house hunting in the neighborhoods around NASA.

In August Al called. ‘‘Don,’’ he said, ‘‘I’m sorry to tell you you weren’t selected.’’ We talked for a few more minutes, and I’m sure he realized my disappointment. They had chosen eleven for the scientist-astronaut class of 1967, including the geophysicist Anthony England, the only other earth scien­tist. I didn’t ask why I wasn’t selected; I was sure he wouldn’t give me any specifics. I rationalized that it was a combination of things. My hiatal hernia (they didn’t have to take any chances on its causing a problem); my seniority (from a government classification standpoint I would have been senior to most of the astronauts selected earlier); my pilot background, which may have been seen as a negative (I would have been the only one they didn’t have to send to pilot training, and that might have made me an apple among all the oranges. What would they do with me during the year the others were in training?) Finally, they might have received some negative comments from MSC man­agers I had disagreed with in years past.

Alan Shepard died recently, so I won’t get a chance to ask him why I wasn’t chosen. Perhaps he would have told me, perhaps not; most probably, after so many years he wouldn’t even have remembered. In any case, the rejection probably did those of us not selected a favor from a career standpoint. Within three years the post-Apollo missions, the prime reason for the selection, were canceled, and none of the class of 1967 flew on a space mission for fifteen years; Joseph Allen was the first from this class to fly as a mission specialist, on shuttle flight STS-4. A few retired or left NASA before taking part in any NASA mis­sions, and several, like Joe and Story Musgrave, made major contributions to NASA programs.

Returning to training and simulations, geological field training for the astro­nauts became more and more realistic and intensive as the date for the first landing came closer. By 1966 all the astronauts had had some level of both classroom and field training. Those in the first three groups selected had the most extensive geological training. Since no one knew who would ultimately be selected for the landing missions, we tried to have them all at as high a level of competence as possible within the time available. Many noted geologists volun­teered to assist in the training; some stayed on to become members of the Apollo Field Geology Team and worked with the astronauts until the last mis­sion, Apollo 17, was safely home. Lee Silver, Richard Jahns, Aaron Waters, Dallas Peck, and William Muehlberger come immediately to mind as volunteers who devoted a significant part of their professional careers to these efforts. Many others made important contributions to astronaut training, including many geologists on the staff at MSC.

I was able to take part in several field geology training trips, and those I attended were all memorable. A specially arranged visit to the Pinacate volcanic fields in Sonora, Mexico, just over the border from Arizona, had a somewhat different purpose. This trip took place in late summer 1966. The Pinacate area includes an interesting set of volcanic craters formed by the explosive release of superheated underground water; craters of this type have their own geologic name—maars. From the air they have an uncanny resemblance to some lunar craters: their rims are only slightly raised, the craters themselves are symmetri­cal, and many are relatively shallow. Some of those at the Pinacate are quite small, a few hundred feet across, and two are very large, the largest being over one mile in diameter. The area where they occur is desolate and isolated, a perfect place to take a high profile group like astronauts, where no one would disturb their training. (It was definitely a place where reporters would not go, for there were no amenities of any kind.)

The Pinacate became one of the favorite training sites, and most of the astronauts made a visit at one time or another. This visit was without astro­nauts; its purpose was to educate my bosses, Phil Culbertson, who had replaced Tom Evans in August 1965, and his boss, E. Z. Gray. Since we were still looking for new training sites for the post-Apollo missions, I thought it was important to show E. Z. and Phil how we would use such sites and what benefits could derive from good terrestrial analogues like the Pinacate. I had arranged with Gene Shoemaker and Al Chidester to conduct the trip as if it were an astronaut training trip, with Phil and E. Z. being treated, in a manner of speaking, as the training subjects.

For both of them it would be a real eye opener; we would camp out in tents for two days in the middle of nowhere, something they had seldom experi­enced. We all flew in to Phoenix and were met by the USGS staffers who would support the trip. Then in a caravan of four or five trucks we turned south on Route 85 with a first stop at Ajo. At that time Ajo was a copper company town with a company store that sold provisions at a discount; the USGS guys always knew how to save a buck. Among other food, we bought frozen T-bone steaks to grill over an open fire the first night; with no refrigeration, we had to cook them that day, and by the time we made camp we expected they would be thawed. From Ajo south, Route 85 takes you through Organ Pipe Cactus National Monument, a unique desert habitat with numerous large saguaro cacti standing like statues along the highway and stretching off into the distance in all direc­tions. This was the ‘‘green desert,’’ with all kinds of unusual plant life including mesquite, palo verde, cholla, and other thorny stands of wicked-looking cactus that I had first seen when we visited Fort Huachuca.

We crossed the border at Lukeville and turned west on Mexico Highway 2. Almost immediately the landscape changed dramatically, becoming much more barren and arid with only a few scattered houses along the road out of Sonoita, the small Mexican town opposite Lukeville. After a few miles we turned off on a dirt road and continued south; the dirt road turned into two tire tracks, and finally we drove in and out of the dry arroyos, gaining a little elevation, and arrived at the volcanic fields about three in the afternoon. While the USGS support team set up camp, we walked over to the rim of Elegante Crater for our first look at the next day’s simulated training site. Elegante Crater is impressive. Over five thousand feet in diameter and eight hundred feet deep, it was not unlike Meteor Crater in many respects, except there were no large blocks of ejecta around the rim and few blocks or large boulders in the interior. The crater looked as though it had been scooped out of the desert by a large spoon, and whatever had been in the center had disappeared. These craters normally constituted a difficult test for the astronauts to interpret and describe so that the accompanying geology staff, acting out the role of a support team back on Earth, could develop a reasonable geologic map based on the astro­nauts’ descriptions.

By the time we returned to camp the tents were all set up and a campfire was lit. Gordon Swann and I went back to the pickup for the frozen steaks and lifted the cardboard carton to carry them over to the cook. They had thawed, the carton had turned to mush, and the thirty or so steaks fell through the bottom onto the sandy soil. What a mess. With a carefully rationed supply of drinking water to last the two days, we could spare only a little to wash off the steaks, so they were still crusted with sand when they finally hit the grill. E. Z. and Phil, along with the rest of the crew, were treated to a new dinner sensation: steak that wore your teeth down if you bothered to chew. I could tell E. Z. wasn’t enjoying his outing—not the best way to impress the bosses with how well organized we were on astronaut training trips. Around the campfire that night the veterans of this type of trip told tales of previous visits to the Pinacate and described some of the exploits they had been party to. Some of the astronauts were enthusiastic card players, and apparently a few exciting card games on past visits had gone on into the wee hours, affecting their next day’s concentration and ability to absorb some rather detailed geological lectures. As we knew, not all the astro­nauts took the field training seriously.

We were a much more sedate group than some in the past, except that a couple of USGS staffers had brought the makings for powerful after-dinner drinks. By the time the storytelling was in full swing, several in the cast were oblivious to the heat and sand. Those of us who were not imbibing heavily decided to call it a day, and along with Phil and E. Z. we crawled into our tents. With fewer seniors around the campfire to dampen the storytelling, the talk grew louder and louder, punctuated from time to time by the equivalent of an Arizona rebel yell. Finally E. Z. couldn’t take the noise any longer. He jumped out of his tent and threatened to cut off all USGS support if they didn’t imme­diately shut up and go to bed. This got their attention; the noise decreased to a low rumble and then silence. When we finally fell asleep, all we could hear was the buzzing of the night insects.

The next morning, up with the sun, we were gathered around the fire awaiting breakfast and the first geology lecture when we noticed that two staff members were missing. We searched around the campsite and couldn’t find the midnight revelers. We were getting worried; rattlesnakes, scorpions, and gray wolves inhabited this area, and there was even an occasional panther. Finally we found one of them asleep in a truck cab, and the other turned up several hundred feet from the camp, lying near a clump of cholla, slightly the worse for wear with his shirt torn and a little bloody. Thus was added another chapter of tall tales for future astronaut training trips. But for E. Z. Gray it was the last straw; he cut his visit short and was taken back to Phoenix that afternoon. By the time I got back to Washington he had calmed down, and we continued to support USGS in all its work. Training trips to the Pinacate were considered highly successful, and on missions to the Moon some of the astronauts would comment on how much the Moon’s surface looked like their memory of the Pinacate.

Mission simulations for crews assigned to specific Apollo lunar landing flights had a somewhat different aspect. For these exercises the two astronauts assigned to the lunar module would be involved, often with their backup crew and sometimes with the command and service module crew member, depend­ing on the objective of the simulation. This meant a support crew of dozens. In addition to the astronauts, lecturers, and technicians, the ever present MSC photographers would be milling around snapping pictures from all angles. Walt Cunningham’s simulation at Bend, Oregon, was an intimate gathering (with the exception of the press that was present) compared with these later simulations. As we approached the flight date, simulations would progress from casual dress at analogue field sites to full suited simulations at MSC or KSC, with some of the latter attempting to follow projected lunar timelines as closely as possible.

As principal investigators were identified for each of the science experi­ments, they would also attend from time to time, along with the contractors building the equipment, so they could observe how the astronauts deployed or operated their instruments. At times the simulations would result in changes to accommodate the astronauts’ ideas on how to improve their interaction with the particular experiment; but whenever possible the astronauts attempted to adjust to the idiosyncrasies of the experiment and achieve the best results for the PIs.

By this point in the training (crews being selected for specific missions), the simulation sites included an MSC high-bay building, the ‘‘back lot’’ at MSC, a small outdoor site at KSC, and a few special analogue sites scattered around the country, chosen to be most like what the astronauts would find on the Moon. The MSC ‘‘back lot’’ or ‘‘rock pile’’ was a few acres of simulated lunar terrain with an LM mockup in the center. The surface was covered with gravel and sand and salted with various types of rocks. A smaller simulated outdoor lunar surface was built at KSC, primarily as a convenience for the astronauts, who spent more and more time there as their launch date approached. The KSC site was often unusable because the ‘‘craters’’ would fill with water at high tide (very unmoonlike), but this site permitted last-minute reviews of specific tasks that may have been added or modified since the previous simulations at MSC. The KSC outdoor site did not include an LM mock-up, so it could support only limited types of simulations. However, there was an indoor site that did include an LM simulator. The KSC simulations were usually conducted in pressure suits to be as authentic as possible. Equipment provided was spare flight article hardware or the closest copy we could obtain.

One of the special analogue sites was near Sunset Crater, a few miles north­east of Flagstaff. Calling it an analogue is a bit of a misnomer, because it was in fact the closest copy of a moonscape that existed anywhere on Earth. Some of the staff at Flagstaff hit on the idea of duplicating the lunar surface as seen in one of Lunar Orbiter’s pictures. They carefully analyzed the selected frame, measuring the diameter and depth of all the small craters and interpreting the history of this small piece of the lunar surface by determining the relative age of each crater based on how the ejecta layers overlay each other. After these cal­culations were made, Norman ‘‘Red’’ Bailey and Hans Ackerman, two Astro – geology staffers, laid out a grid of fertilizer bags on a ten-acre volcanic ash fall south of Sunset Crater. When the fertilizer and fuel oil explosive was detonated, the Orbiter photo was recreated. Not only were the bags arranged according to the explosive force they would generate to create the proper size craters in the correct locations, but they also were timed to go off in the sequence that would provide the correct ejecta layers observed on the real lunar surface. It was a roaring success in all respects, and the creation day was delayed until I was able to witness it on one of my frequent trips to Flagstaff. A movie was made of the explosions, and it was great fun to replay it for visitors who came to watch the astronauts training at the site; each new crater erupted in sequence, in slow motion, and the fine ash flew skyward in great dark jets.

This site, and two additional sites formed in the same manner, became the last tests for the astronauts, requiring them to use all the observational skills they had gained. As they walked or drove around on the closest thing to the Moon they would see until they actually landed there, they described it to the backroom crew so that a geologic map could be made. After completing the exercise, they would review their observations with their instructors to correct any misinterpretations they might have made. All the astronauts from Apollo 13 onward trained at these sites, and I always thought it was one of the best simulations they were involved in, since it was the most complete test of their skills at observation and description.

A drawback with all the pressure suit simulations was that we could not replicate the one-sixth gravity field they would experience on the Moon. In some sessions we tried to simulate the low lunar gravity by using two types of simulators and specially rigged harnesses that partially suspended the test sub­ject and reduced his weight to one-sixth of his Earth weight. These simulations were usually not very satisfactory because the complicated harness setup would reduce only the astronaut’s apparent weight, not the weight of the equipment he was working with. But some of these tests provided important insights, since the mass of the equipment was accurate and the astronauts got a feel for this unique combination of forces. The NASA airplane, normally used to simulate low or zero gravity, also was a poor substitute because of the short duration of each flight parabola. Neutral buoyancy simulations (held in a tank the size of a swimming pool)—a much better way to simulate low gravity environments and the standard way to train for today’s shuttle missions—were in their infancy. They were used for simulating the zero gravity parts of the missions, but not for lunar surface tasks.

In addition to simulating the geologic tasks they would carry out, the astro­nauts simulated the deployment of the Apollo Lunar Surface Experiments Package and the use of all the other equipment and experiments they would carry on the mission. For the final three missions the important equipment additions were the lunar roving vehicle and the lunar drill. The LRV’s deploy­ment from its stowed position on the LM landing stage became a critical part of the timeline. To accomplish all the tasks planned for the extended-staytime missions, the astronauts had to get the LRV functioning as quickly as possible. This meant removing it from the LM stowage bay and setting it on the surface while simultaneously unfolding the wheels tucked beneath the frame, erecting the TV and communication antennas, and finally checking the drive system to be sure it had survived the long journey. A clever but complicated system of cables, springs, and hinges was designed for the LM and LRV.

Once they were sure the LRV was operating correctly, they would load it with the other equipment and experiments that depended on the LRV for their operation. LRV deployment was rehearsed over and over again to reduce the time it took and try to ensure success. During the training sessions the MSC and KSC staffs would introduce hang-ups in the deployment of the LRV and other equipment to see if the astronauts could overcome such adversity. They soon became adept at doing this and foreseeing problems.

Another important task to simulate was getting the loaded lunar sample return containers back into the LM from the lunar surface. This maneuver tested the ingenuity of the MSC engineers because the astronauts could not carry the bulky containers up the LM ladder. They devised a pulley system. One astronaut would kneel in the LM hatch while the other stayed on the surface to hitch the containers to the pulley cables and slowly pull them up to the waiting astronaut. Although it was a relatively straightforward solution, the cable sys­tem tangled easily, so it took many hours of practice to rig the pulleys and coordinate the two astronauts’ actions. Lending urgency to these ‘‘rock box’’ simulations was the knowledge that of all their tasks this was the most impor – tant—the harvest of Moon rocks and soil. If for some reason the sample con­tainers were left behind, the mission would be deemed a failure. This would be especially true for the final missions, which would include samples from loca­tions far from the lunar equator and precious cores collected from below the lunar surface by the lunar drill.

By mid-1967, detailed training and simulation schedules were set up for each of the lunar landing missions.10 Starting forty-four weeks before their scheduled launch date, the astronauts would follow a tight schedule designed to cover all aspects of the missions. Almost 2,200 hours of training and briefings were crammed into their workdays at both MSC and KSC. Some required the presence of all three astronauts, others called for the CSM pilot alone, or just the two Moon-landing astronauts. This constituted a scheduled fifty-hour workweek for each of the three astronauts and the backup crew, with untold extra hours of unscheduled time. They underwent a minimum of 5 hours a week of physical training, 6 hours a week of flying time, 5 hours a week of Apollo flight plan reviews, and 25 hours of flight-suit fit checks, 196 hours of spacecraft tests, 20 hours reviewing stowage procedures for both the CSM and the LM, 40 hours of planetarium exercises to ensure that the crew could use celestial navigation to update their programed navigation system in case of several possible failures, 10 hours of egress training to cover water recovery from the CSM after splashdown, 269 hours of briefings and simulations for science operations, and many other types of training. The 269 hours of science training was one of the largest time allocations, and it was jealously guarded by those of us involved in providing the science payloads, since the other side of the NASA house—the engineers, flight controllers, and other critical partici­pants in launch preparations—would try to preempt some of this time for their own use. But in spite of this constant demand for more astronaut time to attend to nonscience matters, Deke and Al stuck to the schedule, and we were seldom shortchanged. After being named commander for Apollo 14, and while involved firsthand in the training for his mission, Al became a strong supporter for the science team’s training requirements for the final three missions.

When the contract was signed to build the LRV for the last three missions, Rutledge ‘‘Putty’’ Mills, our vehicle guru at Flagstaff, was charged with building a training vehicle that would approximate the LRV configuration so that we could continue to do mission planning and simulations at Flagstaff. (The flight version of the LRV could not be used in terrestrial simulations because it was designed to operate in lunar gravity. It would have collapsed under the astro­nauts’ Earth weight.) An LRV simulator that could be used in Earth’s gravity was not due from the contractor for some months, and we wanted to get an early start on our simulations. Putty did his usual innovative job of construct­ing a vehicle from odds and ends and his fertile imagination. We named it ‘‘Grover the Rover,’’ for one-g rover, and it was ready for testing by the end of June 1970, just six months after Boeing was given the final LRV specifications. At the end of August we conducted a full-scale test, with astronauts participat­ing as well as others. Astronauts in attendance were John Young, Charles Duke, Tony England, Gerald ‘‘Jerry’’ Carr, William Pogue, and Fred Haise.

The test was scheduled to be conducted at the Cinder Lake Crater Field Number 1, but most of the driving over the next four days took place at a vacant lot near the USGS building in Flagstaff. The astronauts present operated the Grover, as did engineers from MSC, MSFC, and NASA headquarters. Putty had built the Grover to run on electric motors like the real LRV, and he had three battery packs available to recharge so we could have more or less continuous operation. At full throttle the Grover could make seven miles an hour carrying two passengers, similar to what we could expect of the LRV on the lunar surface. Mock-ups of some of the tools were stowed on a pallet on the vehicle, the way we anticipated they would be carried on the Moon, although a final stowage configuration for the LRV had not been decided. At the end of the test, all agreed that the Grover would be a valuable addition to future mission simulations, especially when Putty had a chance to add refinements such as a navigation system and additional mock-ups for the lunar communications relay unit, TV, and other equipment the LRV was scheduled to carry.11 Even­tually we obtained a fully functional spare LCRU for our simulations.

A site selected for the simulations conducted toward the end of crew training for the final missions was on the island of Hawaii. Despite the prevailing view that most lunar features were the result of impact processes, all the astronauts had visited Hawaii early in their geologic training to study the wealth of lunar – like features created by the many active or semiactive volcanoes. Simulations for specific missions were a different matter, more like a final exam. We chose several locations on the island to represent geological situations similar to those the crew might encounter on the Moon. Typifying the Hawaiian simulations, the Apollo 17 crew spent the first four days visiting these sites, then had a day of rest. Dallas Peck, a noted volcanologist who had spent a number of years in Hawaii studying the island’s geology, acted as coordinator and principal lec­turer. The final three days were spent at Kahuku, Hualalai, and the volcanic ash wastelands at the crest of Mauna Kea (elevation 13,796 feet), chosen to repre­sent what astronauts Gene Cernan and Jack Schmitt might find at their desig­nated lunar landing site, the Taurus-Littrow Valley.

At Mauna Kea the staff had prepared a series of traverses around the vol­cano’s summit that would approximate those the crew would follow on the lunar surface. Sampling and description stations had been designated at inter­vals replicating as closely as possible the Taurus-Littrow timeline that had al­ready been carefully plotted by the Field Geology Team for the actual mission. All the surface equipment the crew would deploy or operate, except for ALSEP, was transported to the top of the crater, including a simulated version of the LRV. Putty Mills had modified a local jeep to use as a simulated LRV, a cheaper and less sophisticated version of the Grover and other LRV training vehicles. It also avoided the expense of transporting one of these trainers from the main­land to Hawaii. He had removed most of the jeep’s body and engine so that the astronauts were sitting on open seats on the frame and could climb on and off easily. He had also added racks for their tools and sample bags and a mount for their communication antenna, similar to the stowage on the real LRV.

During this training exercise most of us lived in motels on the coast, either in Hilo or in Kailua-Kona, commuting the thirty to forty-five miles a day to the training sites. Some of the USGS staffers lived closer in an army base and kept most of the equipment we would use each day there. Cernan and Schmitt wore street clothes for these simulations; it would have been too costly and time consuming to try to conduct them in pressure suits this far from Houston. To add some mission reality they wore backpacks similar to the portable life – support system, but with battery power only for voice communication back to our simulated Science Support Room out of sight of the traverses.

Bill Muehlberger, the Field Geology Team PI appointed for Apollo 16 and Apollo 17, was in charge of this trip. He brought several members of his team including George Ulrich, Gerry Schaber, and Dale Jackson. Scientist-astronaut Robert Parker was also on hand, since he had been designated mission scientist and the prime capsule communicator during the periods of extravehicular activity. Muehlberger and his team would man the rudimentary SSR, connected to the astronauts only by radio, plotting their progress as they drove around the summit and communicating through Parker as they would during the actual mission. The Field Geology Team, through trial and error on earlier missions, had devised procedures to assist the astronauts if something unexpected hap­pened or to respond to any questions they might have, and these procedures were also practiced.

Those of us not directly involved in the backroom simulation would follow Cernan and Schmitt from a distance as they drove from station to station, making note of how everything fit together—or didn’t, as the case might be. At the end of the exercise, Muehlberger and his team retraced the traverses with Cernan and Schmitt, reviewing how they interpreted their voice reports, cor­recting their map, and then suggesting ways to improve the crew’s descriptions to produce a better interpretation of what they actually saw.

With the first scientist-astronaut geologist in the crew and a highly moti­vated and well-trained commander, we didn’t expect there would be much need for this type of support, but as with all things NASA, we were going to be prepared. All in all, this Hawaii simulation was about as good as we could get in obtaining a high fidelity rehearsal before the real mission was under way.

We conducted one week of intensive, almost uninterrupted training for both

the crew and the Field Geology Team. Apollo 17 would be the last mission, and Muehlberger was determined that it would be the best if he had anything to do with the training and simulations. In just five months it would be the real thing. A final reward for our efforts had become a tradition. On the last night of these trips, a dinner was held at Teshima’s, a lovely Japanese restaurant high on a hill overlooking the ocean, with Mrs. Teshima providing a royal welcome and a special menu. It was a night of storytelling, practical jokes, and reminiscing, a dinner that all who attended will long remember.