The United States Geological Survey. Joins Our Team

At the same time we were conducting our studies at Marshall Space Flight Center, we began to build a strong partnership with the United States Geologi­cal Survey under the direction of Eugene M. Shoemaker at Flagstaff, Arizona. Gene, an outstanding scientist, colleague, and friend, had a major impact on the program. I will be discussing his contributions in future chapters. To a Rocky Moon, by Don E. Wilhelms, provides many details of Shoemaker’s re­markable career; I also recommend this book if you want to read more on Apollo lunar science.1

After leaving Washington in the fall of 1963, Shoemaker returned to Flag­staff, where he had recently moved with his wife, Carolyn, and three small children. He had chosen Flagstaff for his new office location for several reasons. It had a small-town atmosphere, and there were many Moon-like geological features only about an hour’s drive or less to the east. Another plus, although Gene might have denied it, was that Flagstaff was far enough away that he would be left pretty much on his own, undistracted by his superiors in Wash­ington. But the local geology was the real magnet. Meteor Crater, whose origin Gene had helped unravel, was about to become a star in the geological firma­ment, a place all the astronauts would visit and study. He may have thought the Branch of Astrogeology would go quietly about its business, but its notoriety was to grow as its close relationship to the astronauts became known.

Although Gene was in Washington for about two months after my arrival, our paths had not crossed. It soon became clear that he was someone I had to meet. As our contract studies progressed and I learned about his work, it seemed there might be a good match between his interests and my office’s future needs. His staff was already heavily involved in NASA work, including some projects that could contribute directly to our studies. We talked several times on the phone about the direction post-Apollo planning was heading and agreed to meet and see if we could find areas of shared interest.

My first trip to Flagstaff was in March 1964. In those days the best way to get there from Washington was to catch a late afternoon United Airlines flight to Denver and connect with Frontier Airlines for a milk run to Flagstaff. Frontier had recently started operations as a feeder airline connecting many small west­ern towns with larger cities such as Phoenix, Salt Lake City, and Denver. At this time it mostly used the Convair 240, a two-engine propeller plane. As a pas­senger carrier, it offered basic transportation, noisy and drafty. The crew con­sisted of pilot, copilot, and one overworked stewardess attending to the needs of thirty or forty passengers, a few usually sick from the bumpy ride. Since there were frequent stops at cities such as Colorado Springs and Farmington, New Mexico, the plane never reached high altitudes; it flew just high enough to clear any mountain peaks. So you bounced along, buffeted by the thermals that swirled over the mountains below or the clouds above.

On summer trips you dodged thunderheads and lightning all along the flight path and imagined how rough the landscape below would be in a forced landing. By the time you left Denver in the winter it was dark, so all you could see out the small windows were a few lights from the scattered towns below. At some of the small airfields the nearby peaks, unseen in the darkness, towered above the landing approach path. Flagstaff’s airport, cut out of a stand of ponderosa pines, was just a few miles south of town and near one of those towering peaks, Mount Humphrey (12,670 feet). As I walked down the stairs at Flagstaff on that first trip, I inhaled the aroma of the ponderosas, unlike any forest smell I had ever experienced. It was a crystal-clear, cold night with no sky glow from the nearby city. At seven thousand feet, the stars were the brightest I could remember since my days at sea. It was easy to understand why Percival Lowell had established his famous observatory near Flagstaff.

Flagstaff had grown up as a two-industry railroad town, serving lumber and cattle. The main street stretched for several miles along old Route 66 (also U. S. 40), paralleling the railroad tracks. Now it was mostly a tourist town, a stop along the road to the Grand Canyon, about eighty miles to the northwest. The Grand Canyon, like Meteor Crater, would become an astronaut training site. Flagstaff boasted a small college, with a few thousand students at that time, and several motels, small restaurants, and tourist shops, most with a western or Native American motif. East of town were Sunset Crater and other volcanic features, and continuing east you could drive through portions of the Hopi and Navajo Indian reservations and the Painted Desert.

The next morning Donald Elston (Gene’s deputy—his real title was assistant branch chief) picked me up at my motel and drove me to their temporary offices on the grounds of the Museum of Northern Arizona. Gene met me there, dressed in blue jeans, a western shirt, field boots, and bolo tie—the standard uniform for his staff, although a few were not so nattily turned out. My typical Washington uniform of suit, white shirt, tie, and dress shoes drew some wise­cracks, dictating a change of wardrobe for my next visits. Gene’s offices, in several one-story cinder-block buildings, were not imposing. Furniture was rudimentary and looked like army surplus. Some of the more innovative staffers had built bookcases out of packing boxes, and recently Gordon Swann reminded me that when he first arrived in Flagstaff the only extra chair in his small, shared office was a short plank he laid across his wastebasket. In spite of appearances, you could feel the energy and dedication of the staff Gene was putting together; they hadn’t come to Flagstaff for fancy accommodations.

Gene introduced me to those present—mostly young, some of them recent college graduates—and gave me a short tour. Gene had been selected as a coinvestigator for Ranger and the upcoming Surveyor program. Some staffers were busy analyzing the first Ranger close-up pictures, returned only four months earlier, and preparing for the first Surveyor landing. In addition to the Ranger and Surveyor work, his office had the lead in making the lunar pho­togeologic maps that would be influential within a few years in the selection of potential Apollo and post-Apollo landing sites. Most of this latter work, sup­ported by Bob Bryson at NASA headquarters, was being done at the branch’s offices in Menlo Park, California, using the nearby Lick Observatory telescope. Several Flagstaffers commuted to California to work on their assigned quad­rangles; Gene had tried to get as many of his staff as possible involved in the mapping, for training and simply because mapping all the nearside of the Moon was such a big job. Bryson was already upset that the maps were behind schedule. In mid-1964 their commute was shortened to a few miles when NASA, under a program funded by William Brunk of the Office of Space Science and Applications (OSSA), built a thirty-two-inch reflector telescope on Anderson Mesa, just south of Flagstaff, dedicated to providing geologic maps of the Moon and staffed by personnel from USGS. David Dodgen and Elliot Morris were the guiding hands while the observatory was under construction, and it later became Elliot’s small kingdom, supporting many staffers who spent cold nights at the eyepiece to complete their assigned maps.

Although Bryson had warned me he thought Gene was overloaded with ongoing projects, I intended to offer to support some work at Flagstaff if they could take on additional projects. Our meetings went well, and we agreed to work together on post-Apollo mission planning. The topography and geology of the surrounding area would be ideal for testing some of our ideas on con­ducting lunar missions with long staytimes, and it was obvious that Gene and his staff passionately wanted to be involved in exploring the Moon. To alleviate Bryson’s worries, Gene assured me he could hire extra staff for this new work. We shook hands on developing an interagency funding transfer, and I went back to Washington to start the paperwork. Our handshake would lead to almost $1 million a year in cooperative work, with my office covering all aspects of post-Apollo lunar exploration. By the time the Apollo missions were under way, Shoemaker’s team would receive almost $2.5 million a year from NASA to cover its many assignments.

With the paperwork in motion to transfer funding to Flagstaff, Gene began to assemble more staff. He did this with new hires as well as a little Shoemaker ‘‘suasion’’ of USGS personnel at other offices around the country. He had a good nucleus already on site, and to the adventurous recruits this was a mission unparalleled in USGS. A few old hands and a number of younger USGS staff as well as some new hires soon signed up; some reported to the office in Menlo Park, California, to augment the ongoing work there, but most came to Flag­staff. By 1965 Gene had major pieces of many NASA pies: Ranger, Surveyor, Lunar Orbiter, lunar geologic mapping, astronaut training, the job of principal investigator for the first Apollo landing missions, and post-Apollo science plan­ning. At the height of our efforts, in 1968, over 190 USGS staff members and university part-timers were working at several locations in Flagstaff, including offices in a new government complex north of town.

The primary ventures my office funded entailed laying the groundwork to justify the longer-duration post-Apollo missions. This effort soon merged with a need to influence how the Apollo missions themselves would be conducted. With funds beginning to come in from other NASA offices, Gene organized his staff into three offices: Unmanned Lunar Exploration under the direction of John ‘‘Jack’’ McCauley, to cover the ongoing work for Ranger, Surveyor, and Lunar Orbiter; Astrogeologic Studies at Menlo Park under Harold ‘‘Hal’’ Masursky; and Manned Lunar Exploration Studies directed by Don Elston, the last funded primarily by my office.

Our first order of business was to determine what equipment and expe­riments could or should be included on the post-Apollo missions. We incor­porated some of the early results from the MSFC contractor studies as well as the ideas Gene and his staff had begun developing for the Apollo flights. Hand in hand with these studies went the need to define how the astronauts could best accomplish the tasks within the constraints of their space suits and the limitations of their life-support systems. What combination of equipment and procedures would make the most sense from the standpoint of scientific exploration?

In mid-1964 a letter was sent to MSC, over Verne Fryklund’s signature, outlining our need for space suits and support technicians to carry out our planned simulations. It requested an inventory of vacuum chambers where we might test the equipment with suited test subjects. We expected that by 1967 we would want to use vacuum chamber tests to demonstrate that, wherever we were in our studies, equipment design, and procedures, the astronauts could carry out the required tasks. Max Faget’s response about vacuum chambers was encouraging.2 Two large, man-rated chambers, A and B (the larger one ninety feet high and fifty-five feet in diameter) were planned for such simulations. He noted that chamber A could sustain tests lasting several weeks, fitting in nicely with our proposed post-Apollo timeline. We thought Max might be having a change of heart about supporting our needs, since the specifications for the chambers came from his office and the only proposal for such long-duration simulations we were aware of came from us. Until this point there had been no exchange of information between the two organizations, so perhaps Max had paid more attention than we thought to Evans’s earlier briefing.

The situation on space suits was not so encouraging. Borrowing space suits and technicians for simulations away from MSC would be difficult because both were in short supply. Through the intervention of USGS’s Gordon Swann, then stationed at MSC, and others working with the astronauts there, we were able to obtain a surplus Gemini space suit that we trained two staffers at Flagstaff to wear for field simulations. It was not a very satisfactory suit to use in the field, because it was not designed for walking when pressurized, and it was difficult for the wearer to bend at the waist to conduct typical fieldwork. Gemini astronauts either sat in the capsule or, for EVAs, stood almost upright at the end of a tether. But it was useful, especially in the sense that it drove home how difficult it would be for the astronauts, even in a better space suit, to do the equivalent of routine geological fieldwork.

In October 1964 Gordon Swann joined Elston’s group, transferring from his work at Houston teaching geology to the astronauts. Gordon brought his in­sight on how to meet the astronauts’ requirements into everything we were doing, based on his day-to-day interactions with them on their training trips. Gordon soon became our primary suited test subject, pouring gallons of sweat into the boots of our borrowed space suits during his many simulations.

As our studies at Flagstaff accelerated, Elston and his staff began to develop several simulation sites nearby. One of these, just east of town, became a conve­nient place to test our ideas. In July 1964 Bill Henderson and I went to Grum­man to have the model shop build a high-fidelity, full-scale replica of the LEM ascent stage as the starting point for our field simulations. The replica was delivered a few months later. We mounted it on a truck bed, and it was carried back and forth to the field when needed.

With additional help from MSC, we soon graduated to a prototype Apollo suit, which made it much easier to conduct realistic fieldwork, since it incorpo­rated a portable life-support system (PLSS) that let us do away with hoses and hand-carried cooling systems. In June 1965 Gordon Swann and Joseph O’Con­nor were given their first indoctrination into the use of Apollo-type space suits at MSC.3 From that point on, whenever we could obtain the loan of such a suit, we would rehearse and simulate at Flagstaff all the tasks we were planning for the astronauts.

Our simulations and field tests led to the design of various tools and equip­ment to ease sample collection and permit the observation and mapping of geological features. Ideas were tried and rejected and equipment was built and discarded as we learned what would work best. For example, during our field simulations, the USGS “astronauts” practiced viewing the surface from the overhead hatch of the LEM mock-up carried on the back of a truck to obtain, more or less, the correct elevation above ground level. Their experience at taking advantage of this high observation point was passed on to the crews and led to David R. Scott’s decision on Apollo 15 to stand in the overhead hatch to plan his surface activities and traverses at the landing site. Dave Dodgen and

Walter Fahey designed and built a LEM periscope like that recommended earlier for the Martin study (with a few more frills), and it was used successfully during some of the simulations to determine how to study a landing site before the astronauts began their EVAs.

At this point in our work Gene had the good fortune and foresight to bring on board a young geologist who had just finished his graduate work at Har – vard—Harrison H. ‘‘Jack’’ Schmitt. Jack, full of enthusiasm and energy, soon became a leader in our simulation efforts, and with his firsthand involvement in planning post-Apollo missions at Flagstaff, he began his journey toward be­coming (so far) the only professional geologist to walk on the Moon.

We were beginning to make real progress. Not only were we closing in on future tool designs that would work well with a space-suited astronaut, but we were also developing ways for teams back on Earth to process the information that would come back from the Moon in the form of verbal descriptions, experimental data, and perhaps television pictures. At this time a television camera for use on the Moon was not a potential payload item for the Apollo missions. But we believed it would be an invaluable tool for the AES missions, so we usually carried one during our field simulations. We would review the tapes when we returned to the office to complete the analysis of the simulation. We took the next step and set up relay towers on Mount Elden, north of Flagstaff, that let us send the pictures back from the field to an office in the Arizona Bank Building in downtown Flagstaff. After we ironed out the kinks of getting voice and pictures back from the field, we started to design a facility we named Command Data Reception and Analysis (CDRA), where a team of geologists could convert field data in real time into a geologic map. Not only would our planned Moon traverses include geological observations and mea­surements, but we envisioned collecting geophysical information along the route such as gravity and magnetic field measurements. We knew that AES missions would return so much information, collected during miles of traverses by astronauts riding on some type of vehicle, that it would be essential to process the information in near real time. If we could do this, we believed we could redirect the crews or suggest additional surveys to flesh out the picture we were developing of their landing site.

As our CDRA work progressed we brought our ideas to the attention of MSC. This revelation of how we thought the post-Apollo missions should be conducted stirred up a hornets’ nest. We were told in no uncertain terms that the idea would never be approved. Scientists on Earth talking directly to astro­nauts on the Moon? Scientists second-guessing the astronauts on what to do or how to do it? No way! We were told to cease work along these lines. We chose to ignore this ‘‘guidance’’ and continued to improve our vision of how this could be done.

The ALSS-AES missions permitted longer surface staytimes, but to complete the mission and return home the CSM would have to stay in orbit as long as the astronauts were on the Moon’s surface. We began serious study of how we could take advantage of having the CSM in orbit for such a long time. With modifica­tions, in some respects easier to project than extending the LEM staytime, the CSM could remain in orbit for two weeks or longer. What should we do with a CSM that might make three hundred or more orbits of the Moon while the astronauts were on the surface? It seemed obvious: map the Moon from orbit with whatever instruments the CSM could accommodate. In the early stages of these studies we looked at fully automating the CSM sensor package and per­haps converting the LEM to carry three people so that one astronaut would not have to remain alone in orbit on board the CSM but could be on the surface to share the workload. All this appeared possible. We then enlisted the aid of USGS to come up with a conceptual, remote-sensing payload for the CSM. This in turn led to investigating how to tailor the astronauts’ surface activities to provide the ‘‘ground truth’’ that would improve the value of the data returned by orbital sensors. The suite of sensors proposed for the CSM included multi­spectral photography as well as spectrochemical, microwave, and radar instru­ments that would let us extrapolate the data collected at the landing sites to broad regions of the Moon.4

By 1965, three years had passed since the last National Academy of Sciences summer study that led to the Sonett Report. In the intervening time we had learned a lot. Careful study of the close-up views of the lunar surface taken by Ranger increased our confidence that ‘‘normal’’ geological and geophysical studies could be planned for the astronauts. The summer of 1965 was selected as the next date for the Academy to review the status of space science, this time at Woods Hole, near Falmouth, Massachusetts. Dick Allenby and I thought this would be a good opportunity to take advantage of the assembled ‘‘Academy experts’’ such as Harry Hess, Aaron Waters, and Hoover Mackin. I hoped to convene a working group similar to Sonett’s to review our progress and make some specific recommendations for Apollo and post-Apollo science operations.

We made a few calls to see if some of the invited Academy members would agree to extend their time at Woods Hole. Most agreed to stay—it didn’t take much persuasion, since it was such a beautiful spot to be working in the middle of summer. I went to Woods Hole to see if a follow-on meeting could be arranged. In contrast to the twenty participants in the Sonett Ad Hoc Working Group, we envisioned a much larger attendance, probably more than fifty scientists and engineers, including at least one astronaut.

The National Academy of Sciences owned a large mansion directly on the bay at Woods Hole that had been converted to host its many summer con­ferences. With porches on all four sides of the house and broad, well-kept lawns, it was a beautiful, almost idyllic, site. The views of the bay from the conference room windows made you wonder how participants could concen­trate on the business that brought them there. This was my first visit to Woods Hole, and after seeing the mansion I realized that although it could accommo­date the small number of scientists usually invited, it would not serve for the much larger meeting we planned.

A few inquiries turned up no suitable building nearby; we needed a small auditorium for general meetings and several rooms where the various scientific disciplines could meet. Driving around Woods Hole and Falmouth, I noticed the Falmouth High School, a perfect location, and on the spur of the moment went in to talk to the principal (I’ve now forgotten his name). After a brief introduction he gave me a quick tour and said he was willing to ask his school board for permission to host the conference. A few weeks later he called to say it had been approved, and we began the detailed planning for an event that would ultimately involve more than 120 participants.

Developing specific Apollo science guidelines was the first priority of the conference. However, our primary objective for this summer study was to expose the assembled experts to the results of the MSFC contractor studies that we had undertaken for post-Apollo missions. Also, we wanted to show those from the geological community, outside USGS, what we had achieved in more than a year of mission planning and simulation at Flagstaff. During 1964 and 1965 MSC had been steadily adding to its science staff, mostly in the earth sciences, and the frictions I mentioned earlier had been growing. Here was our chance to show them we had received the support of mainstream scientists interested in solving the major lunar problems. Eight of Faget’s staffers were invited, led by William Stoney, John Dornbach, and Elbert King. They partici­pated in two of the working groups and also provided technical advice about telemetry and other capabilities that would be needed to support any proposed lunar science ventures.

Two important attendees were Walter Cunningham and Jack Schmitt: Walt was an astronaut, and Jack was an astronaut-to-be. Jack’s selection in the first scientist astronaut group had just been announced, and his personal involve­ment in our Flagstaff work would be an important step in getting the astronauts to accept our ideas on what to do on the Moon and how to do it. Jack would soon be leaving to start one year of flight training; this conference would be his last official duty as a member of USGS. Walt’s astronaut group, the third se­lected, included many who would become well known, such as Buzz Aldrin and Michael Collins. They had all been given specific Apollo system or technology sectors to monitor and become expert in, besides performing their more ‘‘mun­dane duties’’ of making the transition from military pilot to astronaut. Some had received Gemini mission assignments. Walt’s responsibilities included non­flight experiments, so he was our primary contact in the astronaut corps for any questions about the astronauts’ performing experiments on the Moon. Other astronauts were given this duty as we approached the Apollo launch dates and the more senior astronauts, such as Cunningham, turned their full attention to preparing for specific Apollo missions.

Having Walt at Woods Hole lent immediacy to our planning. Here was someone who might actually carry out our recommendations. Astronauts’ at­tendance at meetings like ours was always appreciated. Requests for them to appear all over the country flooded into NASA. The demand had become so onerous that Alan Shepard and Donald ‘‘Deke’’ Slayton finally set up a ‘‘duty cycle,’’ with each astronaut spending a week or so making public appearances so the others could get their work done. They called this duty being ‘‘in the barrel.’’ Some enjoyed the exposure, some hated it, but all tolerated these distractions, knowing that public relations was part of the job. A separate office was estab­lished at NASA headquarters to ensure that the proper priorities were recog­nized when parceling out this valuable resource. Many requests came from members of Congress, and these were usually put at the top of the list. Although most members supported NASA programs, it was to our advantage to keep them all happy, especially at NASA appropriation times. In any case, Walt was an important addition to our conference, and I assume he was happier meeting with us than on some other public relations assignment.

Walt’s message to us on the first day of the conference, however, was not encouraging. Influenced in part by his training and by his own study and analysis of the preliminary mission timelines, he warned us not to overburden the astronauts with scientific tasks. Housekeeping chores would demand a large percentage of their time on the lunar surface. Such things as recharging the PLSS, the astronauts’ life-support backpack, maintaining work-rest or work – sleep cycles, and monitoring LEM systems—all essential to their safety and health and undertaken in the cramped living space of the LEM—must take priority over science. This was a sobering introduction to lunar science and colored our working groups’ deliberations and corridor talk in the days ahead.

Working groups were established in eight scientific disciplines: geology, geophysics, geodesy-cartography, bioscience, geochemistry, particles and fields, lunar atmosphere measurements, and astronomy. Astronomy was added at the eleventh hour in order to review the preliminary findings of our post-Apollo telescope study and to look beyond Apollo to lunar bases when the Moon could become the site of large astronomical observatories. Such installations might include radio telescopes on the farside where they would be shielded from Earth-made noise. At that time there was no intention to include an astronomy experiment on any of the Apollo missions. One of the members of the astron­omy panel was Karl Henize, then at Northwestern University but destined to be in the scientist-astronaut class of 1967. The other seven working groups, how­ever, were all tasked to review and recommend experiments and operations for the astronauts to carry out on both Apollo and post-Apollo missions, both for two-week staytimes and for lunar bases. The number of attendees (123) ex­ceeded our initial plans, and to ensure that the post-Apollo discussions would be favorably covered, we loaded the attendance with MSFC and USGS staff who had been participating in our studies.

Each working group submitted a report summarizing the results of its delib­erations, and the conference report, compiled by Jay Holmes with the help of many in attendance, was released just before Christmas 1965.5 It immediately supplanted the Sonett Report as the authoritative reference for Apollo and post – Apollo science planning and, as we had hoped, fully endorsed our approach to the post-Apollo missions. In some cases the working groups went far beyond the concepts we had been studying at MSFC and recommended much more complex experiments than we had considered. For example, we reported on the early results of our studies on a hundred-foot drill, and the geology working group recommended developing a drill capable of taking cores at least three hundred meters below the surface in order to penetrate any ejecta layer and reach solid rock. Those of us who had been working on the drill studies realized that achieving such a depth would be a real challenge, and after the con­ference we quickly placed a contract with Bendix to take a first look at how it could be done.

The recommendations of the seven working groups for Apollo experiments are too numerous to list here, and many also pertained to post-Apollo explora­tion, but a few are important in the context of the science payload that was ultimately carried on Apollo. The geology working group listed two primary questions to be answered by the first Apollo landings: What are the composi­tion, structure, and thickness of the Moon’s surficial layer? And what are the composition and the origin of the material underlying this layer? Recognizing that time was the most valuable resource in each mission (reinforced by Walt Cunningham’s presentation), the group gave a lot of effort to recommending tools and procedures that would permit the astronauts to quickly gather the information needed. Even assuming that all the post-Apollo missions we were planning took place, only a tiny fraction of the Moon would ever be visited and studied. Thus it recommended that manned lunar orbiters be scheduled as early as possible, carrying a suite of instruments to acquire lunarwide mapping and remote sensing information on the Moon’s surface composition.

In addition to the geology working group, the geodesy-cartography and geophysics working groups made recommendations dealing with surveying the Moon from lunar orbit. In 1964, under the direction of Peter Badgley, we had begun initial studies of the types of surveys that could be done from an orbiting CSM. We received over one hundred proposals or letters of interest from the scientific community about conducting these investigations, covering all types of surveys from photography to chemical analyses. The Falmouth conference strongly endorsed the need for such investigations.

The deliberations of the geophysics, lunar atmospheres, and particles and fields working groups produced a list of experiments to study the Moon’s subsurface as well as phenomena occurring at or near the surface as a result of interactions with the solar wind or cosmic rays. These interactions were of great interest, since it was difficult or impossible to measure them on Earth because

of the interference of the Earth’s atmosphere and strong magnetic field. For these experiments the Moon could be used as a huge spacecraft floating in free space, on which to mount unique detectors.

The geochemistry-petrology working group also made an important contri­bution to Apollo science. Only two members of the working group were NASA employees at the time (Paul Lowman was one), but all who participated would later become heavily involved in the program either as NASA managers or as sample-return investigators. The working group concentrated on outlining the procedures NASA should follow in selecting the scientists and organizations that would analyze the samples returned by the astronauts; many of their proposals had just been received. It also recommended sampling procedures and container designs for returning the samples in as near pristine condition as possible. Finally the members turned their attention to the design of the Lunar Sample Receiving Laboratory (later shortened to the Lunar Receiving Labora­tory, LRL) where the samples would be quarantined, opened, examined, and sorted for delivery to the laboratories of designated investigators who would then conduct the special analyses they had been selected to do.

Having received the endorsements we were looking for at Falmouth, we charged full speed ahead at Flagstaff to further define potential post-Apollo missions. Based on the emphasis at Falmouth, conserving the astronauts’ time became a major objective of our simulations. We also addressed sample return from these longer missions. The weight allowance for return-to-Earth payloads would be restricted, yet the astronauts would undoubtedly collect many sam­ples during their two-week stay. How could they be sure to bring back the most important ones? We proposed a small sample preparation laboratory that they could use while still on the lunar surface, and one was designed by Joe O’Con­nor, David Dahlem, Gerald Schaber, and Gordon Swann with the help of other USGS staffers. In an undated ‘‘Technical Letter’’ Jerry Schaber reported on the results of one of the field tests, probably conducted sometime in 1966.6

The test confirmed that thin sections of the samples for microscopic study could be prepared in this small laboratory, giving the astronauts, who were receiving some rudimentary training in petrography, a first-order idea of what they had collected. (A thin section is made by sawing rock so thinly that light can be transmitted through the slice, telling a trained geologist its mineralogical composition and something of its history.) On the particular test Schaber reported on, they had included a microscope-television system that permitted simultaneous viewing of the thin sections by both the “astronaut” test subjects and geologists back in the CDRA. As Schaber reported, ‘‘It became apparent during the test that such remote petrographic techniques could furnish a great quantity of information. . . far more than could possibly be returned to Earth in the present LEM vehicle concept. . . . The test results indicated that the thin section image alone could be interpreted with surprising accuracy by the CDRA personnel.’’ (Perhaps a lesson for future Mars explorers, who will certainly face the same problems we were trying to address-how to get the most information back to Earth with a limited return payload.) Instrumentation that we studied as part of such a small portable laboratory included rock-cutting and thin – sectioning equipment, a petrographic microscope, several types of spectrome­ters, a gas chromatograph, and an X-ray diffractometer. This concept was presented a year later at the Santa Cruz summer conference, with the recom­mendation that the images seen in the microscope be beamed back to Earth so that they could be analyzed by experts, thus reducing the time the astronauts spent studying the thin sections.

Our mobility studies at MSFC were providing us with concepts for several types of vehicles that could be carried on the AES missions. In Flagstaff, Rut­ledge ‘‘Putty’’ Mills, with the help of others, translated these ideas into a work­ing model by modifying a truck chassis to carry two test subjects. Once we had this vehicle, which we named Explorer, we planned all our simulations around its use. In 1966 we took delivery of our Cadillac lunar rover, a MOLAB (mobile laboratory) working model that MSFC had built by General Motors, Santa Barbara. It was a Cadillac because this MOLAB model cost $600,000 and had a cab so large that two test subjects could live inside and deploy various geophysi­cal equipment as they drove along, without leaving the cab.

When the MOLAB was delivered to Flagstaff, it created quite a stir. It was an ungainly-looking vehicle with four large, tractor-type wheels supporting a fat, cigar-shaped cab with a rather high center of gravity. Shoemaker, watching it being unloaded from the delivery van and thinking ahead to its use in rugged terrain in the field, declared that the NASA-USGS logos painted on the sides would have to be changed. USGS should appear in large letters on the roof, and NASA should be on the bottom. He was sure that during some future field simulation the MOLAB would roll over, and he wanted any assembled reporters to photograph its ignominious fate with the NASA letters showing as the sponsor and USGS safely out of sight. Gene’s recommendation was not fol­lowed, but his low opinion of the MOLAB test vehicle design was duly reported to MSFC and caused a few red faces. Unfortunately, funding for the AES-lunar base programs was reduced two years after we took delivery of this vehicle, and we had few chances to use it in the field. After a short time it was sent to MSFC, where it was later put on display.

While Gene and his staff were on the front line trying to shape lunar explora­tion, we were dealing with the USGS management back in Washington in the persons of the USGS chief geologists, first with William Pecora then with his successor Harold “Hal” James. Our relationships were always friendly, but although it was clear that they liked this infusion of new money, they never seemed totally comfortable with the assignment. Exploring the Moon didn’t quite fit into the mission of an old-line government agency that had helped open the West a hundred years earlier. This attitude was evident even though at the turn of the century the United States Geological Survey’s first chief geolo­gist, Grove K. Gilbert, had been a pioneer in lunar studies.

Pecora and James, at least publicly, were always strong advocates of working with NASA, and on occasion they would be called on to support lunar explora­tion at congressional hearings or other forums. And certainly the Survey was receiving a lot of favorable publicity from their association with our programs. When the astronauts were covered by the media during geology training trips in some remote corner of the country, there almost always was a USGS staffer identified as lecturing to them. Once the landing missions commenced, USGS contributions became well known, and participants in the field geology experi­ment were in constant demand to discuss the missions. Even the most hard­hearted manager in Washington must have felt some pride at seeing his agency so prominently featured with the country’s new heroes.

Shoemaker was considered a bit of a free spirit within USGS, and all the money he was receiving from NASA, not through his own congressional appro­priation channels, was making him rather independent of his Washington superiors. With his successful creation of the Branch of Astrogeology, Gene decided to relinquish his day-to-day management role and once again reorga­nized by setting up two branches, Astrogeologic Studies under Hal Masursky and Surface Planetary Exploration (SPE) reporting to Alfred H. Chidester. By this time, starting with the first funding transfers in 1961, NASA had trans­ferred almost $14 million to USGS for its various activities, and the action was just beginning to heat up for it to support the Apollo landings. (In all, NASA transferred over $30 million to USGS.)7

With the reorganization, in mid-1967 James sent Arnold Brokaw, a manager with no previous experience in lunar studies, to take charge at Flagstaff and make some further management changes. Brokaw’s appearance altered the dy­namics of our work with SPE, and though we maintained cordial relations with him, we found that the best way to get things done was to work around him and go directly to the staff we had come to know so well over the past three years. The personnel changes made at SPE soon after Brokaw’s arrival put our studies in some disarray. Al Chidester, with whom we had cooperated closely, was transferred and no longer had any role in our work. But with the perseverance and cooperation of Gordon Swann and others, we managed to keep things on track, with our eyes focused on the first landing mission and the hoped-for expansion of our ability to conduct exploration in the post-Apollo era.

By the summer of 1967, with the studies at MSFC and USGS described above under way or completed, we had what I considered to be all the key scientific and operational answers needed to justify more extensive exploration and, eventually, lunar bases. We now felt comfortable providing numbers that would help the scientific community accomplish more productive exploration. Science payloads could be at least 2,500 pounds, including a small vehicle, and the radius of operation at the landing site could be up to five miles. Larger payloads might become available as we continued to learn the full potential of the Apollo hardware; we hoped this would lead to MOLAB missions covering much larger areas on the Moon and establishing lunar bases.

We had a lot of new data to share with the scientific community. NASA headquarters had just announced that it would accept proposals for experi­ments for the Apollo Applications Program (AAP),8 the new name for the post – Apollo program supplanting Apollo Extension System. AAP missions were advertised to begin in 1971 and would include both manned lunar orbit and landing missions, the latter with surface staytimes up to fourteen days. In Will Foster’s office we decided it was time for another summer study to gain more support from scientists for post-Apollo exploration and to encourage them to propose new experiments for the AAP missions. Although the AAP was not yet approved, we thought the announcement was the first step toward its formal recognition, and we wanted to be sure there would be an overwhelming re­sponse of new experiments.

Newell and Foster persuaded Wilmot ‘‘Bill’’ Hess, the newly installed head of the Science and Applications Directorate at MSC, to act as the official host of this conference. The idea was to show the scientific community that under his direction MSC had turned over a new leaf and science would now get the attention it deserved in the Apollo program and any programs that might follow. Until Bill’s arrival, complaints from lunar scientists had been steadily building, and some MSC offices gave the impression that they knew best what science needed to be done and would do it their way. Don’t call us, we’ll call you—maybe. MSC was already managing several Apollo science hardware con­tracts, which added to the concern.

Bill Hess, a physicist, was chief of the Goddard Space Flight Center (GSFC) Theoretical Division when he was asked to transfer to MSC at the end of 1966 to lead a new science directorate. His primary mission at Houston was to reorganize the ongoing science efforts and then evaluate why MSC was held in low esteem by many of the scientists involved in Apollo. A tall, heavy man with a commanding presence, Bill was easygoing but with a touch of steel. He had outstanding scientific credentials and knew NASA politics inside out. We all thought he was the perfect choice for the job. I had come to know him well while he was at GSFC and during the Falmouth summer study, and I knew he would be easy to work with. Perhaps a new day would dawn on our relations with MSC.

Hess had an immediate impact on relations with NASA headquarters. Now, for the first time, we had a senior manager on site who was sympathetic to our concerns and who would return our phone calls, a courtesy seldom extended before his arrival. But he never really became one of the inner circle of MSC managers, and the hoped-for improvements were temporary. One problem was that although he was starting a new directorate, he inherited some of the people from Faget’s office who had been giving us all such a hard time—it isn’t easy to fire or transfer civil servants. In his two short years the climate for science improved, but this was soon reversed by his successor.

The site selected for the 1967 conference was the new University of Califor­nia campus at Santa Cruz. Aaron Waters, a noted geologist and coinvestigator on Shoemaker’s Apollo Field Geology Team, had just joined the staff at Santa Cruz and served as the unofficial host. Over 150 people joined us at Santa Cruz, representing all the geoscience disciplines and including a few astronomers.9 MSFC sent only two observers to the conference, because by this time the decision had been made to manage all Apollo science at MSC, and MSFC quickly phased out of most lunar science studies. Goddard Space Flight Center was well represented, led by Isadore ‘‘Izzy’’ Adler and by Jack Trombka, who had returned to GSFC after his stint at headquarters. They wanted to map the lunar surface extensively from orbit using newly developed sensors. Thirty MSC staffers from various organizations attended, including Faget himself, as well as three astronauts: Deke Slayton, Jack Schmitt, and Curtis Michel (a member of Jack’s 1965 scientist-astronaut class).

The large number of MSC attendees attested to Hess’s new influence and perhaps to the recognition that these summer studies were important in shap­ing lunar science. They came prepared to push their point of view on what science the astronauts should conduct and how it should be done. (I should clarify my criticism of MSC, since it does not apply to the organization as a whole. At this time we were able to work with the MSC science staff, although with difficulty, and Hess’s interest in changing the working relationships with headquarters and the science community was smoothing some of the rough edges. Our relations with other organizations at MSC were usually good, and when I was in Houston I could confide in many friends at MSC or sit down at dinner and discuss the state of NASA.)

As we did at Falmouth, we asked the attendees to think in terms of grand exploration missions, and we had the numbers to allow this. With the newly named Apollo Applications Program would come one of the last attempts at persuading Congress and the administration to continue exploring the Moon after the initial Apollo landings. We hoped that the Santa Cruz conference would stimulate the scientific community to continue supporting lunar explo­ration in spite of growing frustrations with attempting to influence the scien­tific content of Apollo.

Our daily sessions were divided into eight working groups, which reported on their findings at the end of the conference. I attended as secretary of the geology working group, which was led by Gene Shoemaker and Al Chidester (one of Al’s last duties before his transfer) and was dominated by USGS staff and university professors who supported the work we had been conducting at Flagstaff. Major recommendations coming out of this working group included (1) increasing the astronauts’ radius of operation beyond walking range, esti­mated to be five hundred feet, by providing wheeled and flying units; (2) developing a dual-launch capability as soon as possible; (3) creating a sample return payload of four hundred pounds; (4) making the geophysical station flexible so we could react to new opportunities; (5) providing an early manned lunar orbital flight to further map the lunar surface in the visible part of the electromagnetic spectrum and other parts as well; and (6) sequencing orbiter and landing site missions that would include landings at the craters Copernicus and Aristarchus. In general, all the recommendations supported the post – Apollo planning we had undertaken in the past four years.

One of the conference’s recommendations was of special interest to me and others. A second scientist-astronaut selection was under way at the time of the conference, and I was in the final group under consideration. Knowing of the sensitive nature of crew selection and the competition for slots on the landing missions, the working groups tried to be diplomatic when making their recom­mendations for crew training and selection. Also, we hoped that Jack Schmitt would be selected for an early lunar mission, and we did not want to jeopardize his chances by being too aggressive in our advice. The recommendation on astronaut selection and crew training included these words: ‘‘For some of the complicated scientific missions in the later part of the AAP, the Santa Cruz Conference considers that the knowledge and experience of an astronaut who is also a professional field geologist is essential.’’ At the time I hoped they would be to my own benefit during the selection of the next class of scientist-astronauts.

Although the Santa Cruz conference endorsed the need for missions after the scheduled Apollo flights, time was running out for AAP.10 The Santa Cruz attendees, representing many renowned scientists, had proposed important studies on the Moon that were not planned for Apollo. These experiments would require payloads and resources beyond what was anticipated for the Apollo flights. By the time the conference came to a close we knew that NASA budget submittals for fiscal year 1969 would not include funds for missions beyond the already funded Apollo flights. What exquisite timing.

At this point in my government career I had seldom come into contact with the Bureau of the Budget (later named Office of Management and Budget), but in the ensuing years, as a senior official at several agencies, I would frequently meet and argue with its staff members. The original ‘‘faceless bureaucrats,” they had enormous authority and no responsibility. If a program failed or struggled because of arbitrary funding cuts, the agency and program managers would bear the brunt of the failure, not the BOB/OMB staff members who had wielded their red pencils. I don’t recall ever encountering an OMB staffer who had managed a real program; they were blissfully unaware of program com­plexities other than dollars. In spite of this rejection by BOB, we continued to plan for dual-launch missions and extended lunar surface staytimes. We could always hope that the upcoming election might produce an administration more friendly to lunar exploration.

In the fall following the Santa Cruz conference, some major organizational changes took place at NASA headquarters that altered the nature of planning for both the Apollo missions and the missions that might follow the first Apollo landings. With these changes several of us, from various offices, moved to the Apollo Program Office. But before continuing the story of Apollo and post – Apollo science, let’s turn back the calendar to where we left Apollo science at the end of chapter 1.