Category Taking Science to the Moon

On to the Moon:. Science Becomes the Focus

On July 16, 1969, along with a multitude of other sightseers (local Civil Defense officials would later estimate one million), my family and I were on hand to watch the launch of Apollo 11. Our Winnebago camper was parked on the shoulder of U. S. Route 1 about five miles north of Kennedy Space Center and the launch site. We had picked our viewing point the night before, feeling lucky to find a spot so close. It had been a madhouse trying to drive near the Cape; no one seemed to care about following normal rules of the road as cars and campers vied for spots and parked wherever they pleased. Local and state police tried to maintain some order, but it was a hopeless job. In the early morning, as launch time approached, we climbed on the roof of our camper to get an unobstructed view, meanwhile listening on the radio to John ‘‘Jack’’ King, ‘‘the voice of Apollo,’’ count down the final seconds.

Old Glory was flying everywhere, and the crowd was in a party mood. The countdown proceeded smoothly, and at 8:32 a. m. the Saturn rocket lifted off ac­companied by loud cheers and many teary eyes, mine included. Beyond a doubt our hearts went with the crew of Apollo 11. This was the second Saturn У launch I had witnessed, but I still wasn’t prepared for the enormous noise and low – frequency reverberations that reached us, even at this distance, in the minute after the Saturn cleared the launch tower. We watched for several minutes as it disappeared to the east, leaving behind a huge plume of white smoke, then we went inside, finished breakfast, and talked about what we had just seen. My sons, only eight and eleven at the time, still vividly recall the excitement of that morning. I was in a hurry to leave because I was due back in Washington in a few days, but we were forced to wait almost an hour before the traffic jam began to move and we were back on the road. Apollo 11 was on its way to the Moon with the first science payloads that men would place on another body in our solar system. If all went as scheduled, Neil Armstrong and Buzz Aldrin would have the honor of making the first direct, close-up studies of how the Moon’s surface looked and how it felt to walk on the Moon in one-sixth gravity. After the landing and takeoff from the Moon, Mike Collins, the command module pilot, would be waiting in lunar orbit to rendezvous with the lunar module, ready to lower his orbit if the ascent stage did not perform as well as planned.

Four nights after the launch, in anticipation of the landing, the Voice of America (VOA) had assembled a team to report on this once in a lifetime adventure for its worldwide audience. Several NASA colleagues, Merle Waugh, John Hammersmith, William Land, and I, were in the Washington studios as ‘‘color commentators” to back up the VOA reporters led by Rhett Turner, who would be reporting from the Manned Spacecraft Center in Houston. We lis­tened anxiously, just like millions of others around the globe, to the exchange between the capsule communicator (CapCom) Charlie Duke and Armstrong and Aldrin in the Eagle as they went through the final maneuvers to land the LM. The excitement of those last few minutes, heightened by the crew’s diffi­culties in selecting their landing site with alarms ringing in their ears and their fuel supply nearing exhaustion, made Armstrong’s announcement ‘‘Houston, Tranquility Base here, the Eagle has landed,’’ almost anticlimactic. We could hear the cheering in the Mission Control Room through Rhett’s microphone, and we in VOA’s Washington studio were yelling and pounding each other on the back too. Although we had worked for years to help achieve this moment, it seemed incredible that we were successful on the first try.

We were primed to discuss the mission in great detail, but as the night unfolded only a few questions were directed our way, and I was never called on to demonstrate my vast insight into things lunar. VOA wasn’t about to share the limelight on this historic occasion. I did, however, receive a card from some friends in Colombia who said they had heard me on VOA. They told me how proud they were of Apollo 11 ’s success and congratulated me on being part of the program. I wondered if some of my former colleagues remembered their skepticism six years earlier when I decided to leave Mobil and join NASA. I certainly did not regret the decision. Our great hopes to follow Apollo with extensive exploration and lunar bases now seemed remote, but important work still lay ahead to make each succeeding mission more scientifically productive.

As the scheduled launch date for Apollo 11 drew closer, NASA management became more and more cautious and conservative. This was especially evident at MSC, where caution was the trademark, but even at NASA headquarters one could sense growing concern about the many uncertainties and dangers that simulations and planning could not make go away. Mueller’s decision to go to ‘‘all up testing’’ had eliminated several test flights that would have provided additional experience, but it was too late to go back and build confidence any further than where we were in July 1969. The only alternative was to schedule a conservative mission profile leaving as much margin for error as possible.

The Early Apollo Scientific Experiments Package (EASEP) that Armstrong and Aldrin would carry on their flight, described in chapter 7, did not represent a complete Apollo Lunar Surface Experiments Package (ALSEP), since both headquarters and MSC feared that the tasks originally planned would be too demanding. EASEP included a solar-powered seismometer and an additional experiment, the Laser Ranging Retro-Reflector (LRRR). The Swiss-sponsored Solar Wind Composition collector would also be deployed, but its scientific value would be degraded because of the short time it would be exposed to the solar wind. Sample collection and photography were scheduled in connection with the crew’s geological study, but they were also reduced in scope from the original plans.

Before the launch, word of changes had reached Congress, some of whose members were already chafing at the expense of the program. These changes had raised questions about the cost of removing the planned equipment from the Apollo 11 mission. On March 13, 1969, just four months before Apollo 11 ’s scheduled launch, the House Subcommittee on Space Science and Applications held a hearing at which a number of questions were asked about the last-minute science payload changes. Chairman Joseph E. Karth (D-Minn.) asked, ‘‘Can we put in the record why the ALSEP is not flying on the Apollo trip as originally planned?’’1

Our office responded four days later with the following explanation:

The goal of the first Apollo mission to the lunar surface is the successful landing and safe return to Earth of the astronauts. The primary objective of the mission is to prove the Apollo system-launch vehicle, spacecraft, space – suits, men, the tracking network, the operational techniques.

The first landing mission represents a large step from orbital operations.

The descent, landing, extravehicular activity (EVA) and ascent from the lunar surface are new operations in a new environment. Our Gemini EVA experi­ence showed that a methodical increase in task complexity was necessary in order to understand and operate in the zero g space environment. The 1/6 g lunar surface environment will be a new experience. We cannot simulate it completely on Earth. We find, for example, that we simply do not have as much metabolic data as we would like in order to predict with high confi­dence, rates in a 1/6 g environment. Only educated guesses are possible on the difficulties the astronaut will have in maneuvering on the surface or the time it will take him to accomplish assigned tasks.

Until recently, the first mission plan called for two periods on the lunar surface (EVAs). During the second period, the crewmen would deploy the Apollo Lunar Surface Experiments Package (ALSEP). This would take place immediately prior to lunar ascent and rendezvous. Because of biomedical unknowns, we are concerned with the degree to which the second EVA would fatigue the crew and adversely affect their performance during the critical ascent and rendezvous phases of the mission.

After extensive review and evaluation, we reached the decision not to have a second EVA on the first landing mission. The ALSEP will be deferred to the second mission. We will make every effort on the first landing to obtain data leading to a firm assessment of the astronaut’s capabilities and limitations on the lunar surface with a view toward increasing, on subsequent landings, the percentage of EVA time available for scientific investigations. Deployment of the ALSEP on the second mission is planned as a primary objective.

Our answers to other questions raised by the subcommittee included an esti­mate of $5 million to modify the ALSEP seismometer to the EASEP configura­tion. (This number differs from the contract cost of $3.7 million discussed in chapter 7 because it includes other costs associated with the EASEP, such as integration and training, that were not part of the Bendix contract.)

Left out of the response was another concern, the performance of the LM during the first landing and takeoff on the Moon. Although the LM had per­formed well on Apollo 9’s Earth orbital flight and Apollo 10’s close approach to the Moon’s surface, leaks in its propellant tank had only recently been fixed. With only two LM test flights under our belts, NASA management was still concerned about this problem. Our office was understandably chagrined at the

changes in the timeline and the science payload, but this turn of events lent even greater importance to ensuring that the science planned for the next landings was not compromised.

Another interesting exchange before a Senate committee took place shortly after the House subcommittee hearings. Homer Newell and John Naugle ap­peared before the Senate Committee on Aeronautical and Space Sciences on May 1. During the questioning, Senator Carl T. Curtis (R-Neb.) asked Newell and Naugle if knowledge gained from our completed space missions had changed previous beliefs. Both Newell and Naugle said yes, and Newell went on to provide a surprising example. He said that the “mascons” discovered by tracking the Lunar Orbiter flight paths (concentrations of high density material below the surface of the lunar seas that might indicate large meteor impacts) ‘‘give rise to some of the speculation that maybe at one time these areas were actually oceans or seas and [that] sediments from these oceans or seas is what filled those holes.’’ You won’t find these speculations in chapter 2, although many thought there was a chance that some water had been present on the Moon at one time. The theory that the Moon once had oceans was not sup­ported by any prominent theorists of the day, and if the large impact craters had been filled with sediment of some kind they would have been deficiencies of mass, not mass concentrations. For a crater to be a ‘‘mascon,’’ the fill had to be some unusually dense material. Even senior NASA managers had a hard time keeping up with changing theories as new information was gathered and ana­lyzed by more and more students of the Moon.

One week before Apollo 11 lifted off, Sam Phillips issued a new Apollo program directive (APD) detailing a total of ten lunar landing missions.2 The first landing was designated a G mission with the characteristics noted above, and the next four were called H missions. The H missions were designed around two EVAs, surface staytimes of up to thirty-two hours, and our old reliable payload of some 250 pounds. The final five missions, Apollo 16 through Apollo 20, were called J missions. Although the APD did not specify any science payload numbers, it stated that both the lunar module and the command and service module would be improved to permit longer staytimes. We anticipated that the LM would be able to carry additional descent propellant, which would translate in part to an ability to carry larger science payloads. We still held out hope in 1969 for flights beyond Apollo 20, but realistically we would have to extract as much science as possible from these ten missions. It was not exactly what we had planned for in 1964 and 1965, but we expected the J missions to be far better than the original Apollo plans. An interesting statement in the APD was that the constant-volume space suit would be available for the J missions. This never came to pass, and if such suits had been used they probably would have had little effect on the productivity of the J mission EVAs. LM and CSM consumables became the limiting factors, not the astronauts’ metabolic rates.

As we had simulated at Martin Marietta in 1964 and 1965 in case of an abort after touchdown, the crew of Apollo 11 first used their eyes to describe the lunar scene and took a few photographs before leaving the LM. One other piece of data collected was a movie of the landing site filmed from Aldrin’s window as Armstrong maneuvered for the landing. Not much scientific use was made of this movie because of its limited view of the surface, but you could see how the Moon’s surface layer was disturbed by the exhaust of the LM descent engine, with the fine-grained particles shooting rapidly away from below the LM in a fuzzy blur. These pictures confirmed that the lunar surface reacted as predicted to the LM exhaust and helped ease concerns about future LM landings. Peering out his small window, Armstrong provided the first descriptions of the surface, and Armstrong and Aldrin took pictures with the Hasselblad camera and de­scribed what they could see from their windows. Their observations added to the overall understanding of the landing site but did not reveal precisely where they had landed.

Whether Armstrong or Aldrin would have the honor of being the first human to stand on another celestial body had been decided long before Apollo 11 was launched. The initial timelines, circulated almost a year earlier, had indicated that Aldrin would be the first out. As planning for the mission ma­tured, however, it became evident that the LM commander, Armstrong, would be in the best position inside the LM to perform this historic first, seniority notwithstanding. From a science standpoint it really didn’t make any difference who would be first on the surface, but for Aldrin the decision was obviously a disappointment, and it continued to trouble him years later. Usually few people remember who was the second to do something; however, both Armstrong’s and Aldrin’s names are synonymous with the first Moon landing. Through the years Aldrin has received his deserved recognition, but he is not quite as famous as if he had been the first to touch the Moon.

It took some time for Armstrong and Aldrin to secure the LM and get it ready for a quick takeoff, if necessary. After landing and preparing for an emergency takeoff, the timeline scheduled a meal followed by a sleep period. The astro­nauts, understandably excited and not sure how long they would be permitted to stay on the Moon, asked Mission Control to skip the sleep period and immediately begin preparing for their EVA. Receiving approval, they donned their space suits, and a little under seven hours after they landed Aldrin opened the hatch. Armstrong squeezed through and bounced down the ladder (without seeing any exploding ‘‘Gold dust’’).

His descent and first steps on the Moon were recorded for all the world by a television camera attached to the landing stage, which he activated from the top of the ladder. This camera, built by the Westinghouse Aerospace Division, had been the subject of much debate. Could we afford the weight (about ten pounds) and the complications of deployment, since we knew the quality of the pictures would be poor? I was for not carrying it, especially when we were discussing whether to include the ALSEP because of weight and EVA time concerns. But once it was decided to eliminate the ALSEP, the question became moot from a science perspective. The ‘‘let’s carry it’’ side won the day, and it turned out to be a valuable tool both for public relations and for science. We used the TV pictures, in spite of their poor resolution, to help reconstruct the astronauts’ movements and plot the geology. Some senior NASA managers complained during the mission about the poor quality of the pictures, but by then it was too late. (The poor picture quality was caused not by any Westing- house design deficiencies but by the NASA specifications, dictated by weight and power constraints and antenna performance.)

After examining the LM and reporting its status, Armstrong began describ­ing the scene around him and his impressions of the lunar surface. Then he took a few photographs. He collected the contingency sample and put it in a pocket of his space suit, and he was soon joined by Aldrin to complete their carefully choreographed EVA timeline. At this point Armstrong removed the TV camera from the LM and set it up some sixty feet to the northwest, provid­ing a limited view of the landing site and of the astronauts’ movements as they went about their EVA. From this time until Armstrong and Aldrin reentered the LM, they performed all their tasks as planned. I won’t go into detail on what they accomplished; references listed in the bibliography describe these activities in great detail. Both astronauts performed all their scientific assignments better than expected under extraordinary conditions. One might think that the first

men to land on the Moon might not have their minds completely on the scientific tasks before them. One might expect them to be thinking about the upcoming liftoff, a maneuver never before attempted, which their survival depended on. Armstrong and Aldrin seemed to put such concerns out of their minds. They appeared to be completely absorbed in deploying the experiments, sampling, and describing what they were seeing and doing.

Aldrin placed the EASEP, the last-minute replacement for the ALSEP, on the surface about sixty-five feet south of the LM and in the same general area as the Laser Ranging Retro-Reflector. He had no trouble unfolding the solar panels and erecting the radio antenna, and once set up the experiment turned on automatically. Back on Earth, signals were received almost immediately, relayed to Houston from the NASA Manned Space Flight Network. We knew it was working because the seismometer recorded Aldrin’s footsteps as he walked nearby, but we hadn’t expected to receive so many signals.

The MSC and Bendix engineers manning the EASEP console in the Science Support Room (SSR) soon began to see a problem. The temperature of the seis­mometer package was rising faster than expected. It took some time to arrive at a probable cause, but they finally decided that dirt and dust were covering some of the surface, reducing its ability to reflect heat. Both Armstrong and Aldrin had commented on how far the soil would fly when they walked, as well as on how dirty their suits got during the EVA. While deploying the EASEP they had completely circled the experiments, so it was logical that some soil had coated the surfaces. Also, based on Aldrin’s comments, as we continued to track the rising temperature after their takeoff, it appeared he had placed the EASEP experiments closer to the LM than requested. We assumed that dust thrown up during the takeoff had also been deposited on the experiment surfaces. We kept our fingers crossed that the soil would not overheat the seismometer and had not obscured the small corner reflectors of the LRRR, making it difficult to bounce laser beams back to Earth.

These eventualities didn’t come to pass; the seismometer survived the rest of that lunar day (fourteen Earth days) and the following lunar night and came back on line for seven more days when the solar panels saw the sun again. The seismometer recorded several interesting events during its short lifetime, in­cluding the shocks of the astronauts’ backpacks hitting the lunar surface when they were thrown from the LM and the small ‘‘moonquake’’ when the ascent stage lifted off. Based on this performance, we could anticipate that the seis­mometers of the same design scheduled for the full ALSEP deployments would provide even more information during their much longer lifetimes.

In addition to still photographs, movies, and the Solar Wind Composition collector foil, a total of some forty-seven pounds of individual rocks, soil, drive – tube cores, and the contingency sample, all neatly packaged, finally found their way to MSC, where the staff at the Lunar Receiving Laboratory, and eventually the sample analysis principal investigators, eagerly awaited them. On the recov­ery aircraft carrier, the USS Hornet, the samples were divided into two batches and flown to Ellington Air Force Base in separate aircraft to ensure that some samples would survive in case one plane was lost at sea. There was always the chance we might not get back again to collect more samples. From Ellington, they were carried to the LRL.

The astronauts, wearing isolation garments that they donned in the CSM while awaiting recovery and transport to the Hornet’s deck, were immediately sequestered in a specially designed trailer lest they contaminate those around them with some deadly unknown virus. After the Hornet arrived at Hawaii, they too were flown back to MSC in their trailer along with two volunteer MSC doctors, to begin their one-month quarantine.

The samples, which had arrived before the astronauts, were carefully opened in the LRL, inventoried, and briefly described. In the meantime we were moni­toring the signals sent back by the passive seismic experiment and attempting to find the LRRR that the astronauts had left behind. This latter operation was not as easy as we expected, since the exact location of the landing site was not immediately known. Mike Collins had attempted unsuccessfully to locate the LM from orbit using the command module sextant. After analyzing the flight data and the returned photographs, we passed our best estimate to the LRRR PIs, and the LRRR was found on August 1, 1969, by the Lick Observatory in California.

On August 23, 1969, one month after Apollo 11 splashed down and the date when the astronauts were released from quarantine, George Mueller forwarded a memo to Clare Farley, James Webb’s executive officer, to be included in the report being sent to the president summarizing the results of man’s first foray to the Moon.3 In his memo, drafted in part by our office, he described the initial scientific results of Apollo 11 and summarized the program adjustments that would be made as a result of the mission. Included with the memo was a preliminary traverse map compiled by Gerry Schaber and Ray Batson of the United States Geological Survey using tapes from the lunar module’s television camera, photographs taken by the astronauts, and educated guesses based on what the astronauts reported from the Moon. The map sent to the White House had been further updated during the astronauts’ debriefings while they were still in quarantine. By this time photographs of the astronauts on the Moon and a few photographs of ‘‘Moon rocks’’ had circulated in all the newspapers and some magazines, so Mueller didn’t include any photographs of the astronauts with his memo, but he did include a photo of one of the returned samples. The Schaber-Batson map had just been completed and represented new informa­tion not yet made public, tying together everything the astronauts had done during their brief stay.

Short and to the point (five pages plus attachments), Mueller’s memo pro­vided an initial age dating of one sample (3.1 billion years), compared the chemical and mineralogical content of a few samples with that of the Earth, and offered the conclusion that the Earth and the Moon probably were formed ‘‘from the same whirling cloud’’ some 4.5 billion years ago. (It wasn’t clear where that comparison came from, but it wasn’t too bad a description if you agreed with the conclusion.) He also briefly discussed some results from the passive seismometer and LRRR; the latter experiment permitted the measure­ment of the Earth-Moon distance to within twelve feet as opposed to the best previous accuracy of about two thousand feet. (The accuracy of a few inches predicted in chapter 7 would come only after several years of ranging from three or more stations.) The last sentence we added to the memo was, we hoped, a thinly veiled plea to the White House to the keep missions going: ‘‘The indications thus far are that the Moon is a celestial body with complex structure, geology, and chemical history that may take considerable effort to unravel.’’

Mueller’s attachment summarizing planned program adjustments had an important effect on the subsequent missions. With the lunar landing mandate successfully completed, Mueller now proposed to slow the pace of the missions from one launch every two and a half months to one every four months. He stated that this not only would save money but would allow us to ‘‘increase mission flexibility and scientific return in later missions.’’ This was a welcome change to those of us planning the science and to the staffs at MSC and KSC, who had been working around the clock to support the shorter schedule. This would, we hoped, allow us to factor in some of the results of the previous missions while developing the objectives for each succeeding one and to alter the science payload and astronaut training accordingly. To a large degree we were able to do this on the last three J missions.

With the flight of Apollo 11 successfully concluded, General Phillips relin­quished his position as Apollo program director and returned to the Air Force. He was replaced by Rocco Petrone, who until this new assignment had been director of launch operations at KSC. Rocco, a West Point graduate, was a large man. He had been a backup tackle on two of Coach Red Blaik’s most famous Army football teams of the 1940s, which won thirty straight games before being defeated by Columbia in 1947, my freshman year. The teams featured ‘‘Doc’’ Blanchard charging up the middle or Glen Davis scampering around the end, at times behind the broad back of Rocco Petrone. He was listed in the game programs of the time as six feet one and 202 pounds; in the 1940s these were not intimidating numbers for a tackle, but he wasn’t exactly small. In 1969 he was a little more imposing, perhaps with a few more pounds than he carried in his playing days.

I don’t have many recollections of specific meetings with Sam Phillips, but I do remember calm, quiet, efficient status reviews that moved along quickly, with Phillips clearly in command—a management style much like George Mueller’s. Meetings with Rocco were different. He came to Washington with a reputation as a hard-nosed, hard-driving manager with his record at KSC—all Saturns launched successfully—a testimony to his management skills and his team’s ability. He had succeeded in what must have been a difficult environment under the early tutelage of the German-trained rocket scientists assembled by Wernher von Braun and KSC director Kurt Debus, both known to be sticklers for detail and perfect performance.

Rocco was the only senior manager I worked with who truly had a pho­tographic memory. If you gave him a ‘‘fact’’ related to your program during a briefing, woe unto you if you changed anything a week, month, or year later. Rocco would catch or challenge you, and he was almost always right. Rocco’s meetings were a little more lively than Phillips’s, especially if there were discus­sions of delays or unexpected changes. He was never shy about showing his displeasure, and it was reinforced by his imposing frame. Conference calls between Rocco and the NASA centers were always interesting. Usually they were arranged to discuss some critical problem, so by their very nature they were bound to be contentious. As we listened to Rocco asking questions in his distinctive high-pitched, singsong voice, we could visualize the speakers at the other end of the line squirming as they tried to justify some earlier position that he didn’t agree with. Rocco soon became our strong right arm and a defender of lunar science. Once he was convinced of the correctness of a scientific position, we seldom lost any ensuing argument with MSC. After Rocco’s arrival we really buckled down to expand and improve the science on the last three missions.

Flight readiness reviews (FRRs) were another area where Rocco ran a taut ship. Hosted by KSC, they were the final review, held about one week before a scheduled launch. Chaired by Chester ‘‘Chet’’ Lee, Rocco’s Apollo mission di­rector, they usually lasted one full day. There were representatives from all the NASA centers involved in the launch as well as the contractors and the required Department of Defense participants—a cast of hundreds. Every aspect of the mission from prelaunch preparation to splashdown and recovery was discussed in detail and checked off as being ready if it passed the rigorous review. Action items or deficiencies recorded during earlier mission reviews were carefully analyzed to be sure they had been properly attended to. This process might result in long debates, followed by documentation to prove problems had been resolved. Any items still open after the FRR were subject to a final review and structured sign-off before launch. Here is where Rocco’s photographic memory was put to the test. He would recall the smallest detail and ask penetrating questions. If the presenter could not answer to his satisfaction, someone had to leave the room and gather the missing information.

FRR attendance was carefully controlled. NASA senior management was seated at the front of the room, along with at least one of the astronauts who would be on the crew or serve as backup crew for the launch under review. Briefers with their supporters scurried in and out as called for by the agenda. For the J missions, I was entitled to a seat at the back of the room to take notes and perhaps pass on a discreet question for Chet Lee or Lee Scherer to ask. But the FRRs tended to be a one-man show, with Rocco calling the shots and the other senior managers like James McDivitt, Deke Slayton, and Al Shepard recognizing his mastery of the occasion. Everyone knew Rocco’s reputation for detail, and no facts or concerns were held back. We all understood that the lives of the astronauts seated in the room with us could be in jeopardy if the smallest problem went undetected or unsolved.

Hangar S became a kind of science headquarters at KSC as we approached the Apollo lunar missions. It was a little seedy looking on the outside—the paint was peeling and the large S was barely readable—but the inside was a high-tech workshop. As the name indicated, it was formerly a hangar at Cape Canaveral Air Station, but it now functioned as an important facility at KSC where final preparations and checks were carried out for all the experiments. Mock-ups of the LM and CSM were maintained in the hangar and used for stowage checks and simulations, which became increasingly complex for the missions follow­ing Apollo 11. The crews would spend more and more time at KSC as they neared the launch date, so it was important to have a place where they could stay up to date on any changes that might involve the experiments.

Flight experiments were sent to KSC from contractors around the country. KSC engineers would receive the flight hardware and store it in a clean room in another building near hangar S where final checks would be made to ensure that nothing had been damaged during shipping. Contractors building the experiments and equipment did their own inspections before the items left their plants, but the final checks were done at KSC. Nothing was loaded on the LM or CSM if it had not undergone a rigorous preflight inspection. Once it passed this inspection, it would be taken to the Vertical Assembly Building to be stowed.

Since ALSEP was the major science payload after the flight of Apollo 11, it received the most attention. It was carefully unpacked in the clean room, and each experiment was set up to check cable connections and any unique fas­teners, thermal blankets, or other apparatus that might give the astronauts trouble during lunar deployment. Chuck Weatherred, the Bendix ALSEP man­ager, recalled an important exchange as he helped the KSC team prepare for the launch of a ‘‘dummy’’ ALSEP on Apollo 10, scheduled to fly to the Moon but not land. Peter Conrad and Richard Gordon, the Apollo 12 crew, came into the clean room to watch the processing of the package that would simulate the weight and center of gravity of the ALSEP so that the MSC flight dynamacists could calculate how the spacecraft would react to various maneuvers during the mission. Although they had visited Bendix and seen their ALSEP in the final stages of manufacture, they knew their training schedules did not call for them to have any direct interaction with it until they were on the lunar surface. Conrad asked Chuck if they could participate in the final checkout before their ALSEP was stowed for the journey to the Moon. Chuck thought that was a great idea and said he would get approval from MSC, but Don Wiseman, his MSC contract manager, turned the request down. MSC didn’t want the astronauts fooling with the flight hardware before they deployed it on the Moon.

After several appeals and backing by the astronauts, that decision was re­versed, and all crews starting with Apollo 11 were permitted to work with the flight hardware at KSC before it was finally stowed for the trip to the Moon. It was perhaps a small victory, but I feel sure it made the crews more confident that they would not confront any surprises. ‘‘Murphy’s Law’’ says anything that can go wrong will go wrong. No matter how closely you monitor the manufac­ture of such a complex set of equipment as ALSEP, minor changes not reflected in the simulation hardware or documentation (someone’s last-minute bright idea) can creep into the design and could cause complications 238,000 miles away. We had few such problems with the science payloads, in part because we worked hard to be sure the astronauts were always in the loop.

At the same time that we were savoring the success of Apollo 11, the National Academy of Sciences’ Space Science Board was conducting another summer study, once again at Woods Hole. The study was chaired by Harry Hess of Princeton University, who had also led the 1965 summer study held in conjunc­tion with the Falmouth conference. Harry was a strong advocate of manned and unmanned lunar exploration, and his position at the Academy as well as his overall reputation in the scientific community lent great weight to our Apollo science planning. Harry’s objective for the study was to capitalize on Apollo 11 ’s success and lend support to those of us arguing with the administration and Congress to use the remaining Apollo hardware to carry out more missions and missions with ever increasing exploration potential.

Immediately after Apollo 11 ’s return, some leading decision makers in and out of Congress, who will remain unnamed, had been quick to propose ending lunar exploration and spending the money saved on various social programs back on Earth. These discordant voices motivated Hess to quickly call for the study. I attended the meeting with Don Wise, who had joined our office from Franklin and Marshall University to be Lee Scherer’s deputy. We made several presentations based on our ongoing efforts for the J missions, pointing out the potential for enhancing the science return. We also reviewed the recommenda­tions of the Santa Cruz conference and the ‘‘Lunar Exploration Plan’’ we had disseminated at the end of 1968. This summer study provided a new oppor­tunity to resurrect some of our old plans for long-duration missions that we were forced to abandon in 1968 for lack of interest by Congress and the admin­istration. Along with many other participants in the Apollo program, I strongly supported Harry’s views that we must make the case to take advantage of this opportunity—to squeeze as much science as possible from the Apollo program. After all, the major expenditures had already been made; using all the hard­ware, and doing it more efficiently, would entail adding only a small fraction to the total spent to date for the new science payloads and mission operations.

Tragedy struck the study on the first day, August 25, 1969. During the morning coffee break Harry complained of chest pains and left to see a doctor. He never returned. We were told he died peacefully at the doctor’s office. This, of course, spread a pall over our meeting. We had lost an irreplaceable leader whose vision had been, since the earliest days, a major force in our efforts to bring good science to the Apollo program. Only a few special people, including Ralph Baldwin, Harold Urey, and Gene Shoemaker, can lay claim to being fathers of lunar exploration, and Harry Hess belongs in that company. We continued our deliberations under a new chairman, Bill Rubey, the newly ap­pointed director of the Lunar Science Institute, and then issued our report.4 A case was made to support the launch of the nine missions still being planned at the time and to continue additional missions through 1975. The study con­cluded: ‘‘The decision concerning the nature of the lunar exploration program after the mid-1970s will hinge on the national commitment to manned space flight and on the significance of the scientific discoveries that emerge in the next few years.’’

While this report was in press, those of us advocating more Apollo science received another blow. Bill Hess resigned from his position as director of sci­ence and applications at MSC; he finally got tired of bucking the entrenched antiscience interests there. Tony Calio, who had earlier worked with us on Foster’s staff, took Hess’s place. When Tony left our office to go to MSC, we gave him a going-away party in Washington, wished him success in taking on such a difficult position, and looked forward to having someone at MSC who would be receptive to our interests. At the time, we didn’t know his appointment would adversely affect our relationship with MSC, but within weeks it became apparent. Tony quickly adopted the MSC line, and our relationship with MSC regressed to where it had been two years earlier. He became hard to reach by phone, and when we did get through he ignored most of our suggestions. He also developed an intense dislike for the staff at USGS. I never fully understood the reason for this antagonism—perhaps it was a holdover of earlier disputes between USGS and some of the staff he inherited. But this undermined USGS’s ability to support the upcoming missions for which members of the Field Geology Team had an ever increasing responsibility. It was only through their close relationship with the astronauts and others in the astronaut office that they were able to influence the geology content of the missions.

Returning to the remaining missions, Apollo 11’s success and a ringing endorsement from the National Academy of Sciences energized many in the science community to propose exciting new experiments for the remaining missions as we geared up to take advantage of a relaxation in some of the mission constraints. Until Apollo 11 returned safely, every Apollo engineer and system and subsystem manager was holding a little in reserve just in case it was needed. A little extra weight, a little extra available propulsion, a little extra performance margin. Slowly, with the help of the Bellcommers, these margins were identified and translated into increased science payload and more operat­ing flexibility.

The Schaber-Batson map was the first attempt, other than during simula­tions, to reconstruct in near real time what was happening on the Moon. Although during the Apollo 11 mission there was no direct exchange between scientists on Earth and the astronauts, based on our Flagstaff simulations we could see how this could be done effectively for the later missions. For Apollo 12 and the remaining four missions we tracked the astronauts in real time and had an up-to-the-minute map of their progress in the SSR. We coordinated our tracking with the flight controllers and medical staff monitoring the astronauts’ performance to ensure that their traverses would not overextend their life – support expendables. This monitoring was especially valuable during the last three missions, when the astronauts were often far from the LM and we had to be sure they had enough life support reserve to walk back if the lunar roving vehicle failed. For the science team it had another important aspect: it allowed us to relay suggestions for modifying the astronauts’ activities through the CapCom as they reported their findings and, at times, changed the timelines on their own initiative.

In September 1969 we advertised the opportunity to propose new experi­ments for the J missions that would utilize the LRV and the longer staytimes. This announcement, while directed primarily to missions 16 through 20, indi­cated that proposals to perform simple experiments on flights earlier than Apollo 16 would also be accepted.5 Perhaps the most ambitious aspect of this announcement was our optimism about where we would be permitted to target landing sites for the flights that would follow the initial landings. Scientifically exciting sites recommended by the Group for Lunar Exploration Planning (GLEP), such as the central peaks of Copernicus and the rim of Tycho, were included as candidates in the announcement so that proposers could consider their unique characteristics for their experiments.

With the arrival of Tony Calio and the immediate change (for the worse) in climate at MSC in regard to science, we began to lobby Rocco Petrone to push MSC to modify management’s responsibilities for science in the hopes that this would improve our working relationship. He talked to Jim McDivitt about making some changes. At the end of October 1969 our office originated a memo for Petrone’s signature formalizing these suggestions. The opening sen­tence, underlined, stated, ‘‘I think we have a problem in the management of the science program which warrants immediate action.”6

McDivitt responded two weeks later and gave us half a loaf.7 He moved the design, development, testing, and delivery of approved Apollo experiments from Calio’s office, the Science and Applications Directorate, to the Engineer­ing and Development Directorate, managed by Max Faget. (We weren’t sure if this was a victory.) S&AD would still be in charge of the scientific requirements, science mission operations, postflight data analysis, and interactions with the PIs, but McDivitt promised that his office would strengthen its science over­sight. This was encouraging, since Petrone and McDivitt usually agreed on the important aspects of the missions, and science would take center stage for the remaining flights. In spite of these changes, our concerns would soon be echoed by the scientific community.

Through 1970, we were still hoping dual-launch missions might be rein­stated, enabling fourteen-day stays on the Moon, and the trade journals of the day continued to write about these plans as if they were approved.8 In Lee Scherer’s office we continued to study an LM shelter and a dual-mode (manned and automated) roving vehicle. Scherer urged Marshall Space Flight Center to complete the preliminary design and promised funding for this work.9 Mean­while, preparations continued for the next landing. Apollo 12, we hoped, would allow us to accomplish some of the science originally scheduled for Apollo 11 but at a different mare site, many miles to the west.

Apollo 12 was successfully launched in November 1969 and landed about eight hundred miles west of Tranquility Base at the lunar feature called the

Ocean of Storms, another mare site. If our photo interpretations were correct and the landing site was on an ejecta ray from the crater Copernicus, a few hundred miles to the north, we hoped to return samples of material from deep within the Moon, excavated by the impact that formed this huge crater, some forty miles in diameter. Copernicus is one of the craters you could identify under proper lighting conditions with your ten-power binoculars, just a little west-northwest of the center of the Moon.

Two EVAs were scheduled and carried out, and a full ALSEP was deployed. Peter Conrad and Alan Bean proved to be enthusiastic lunar explorers. Much was made in the press of Pete’s laughing, giggling, ‘‘cackling,’’ and joking as he went about his tasks, but he and Al performed flawlessly, bringing back some stunning pictures and a wide assortment of lunar rocks. The TV camera, simi­lar to the one carried on Apollo 11, was damaged soon after they climbed down from the LM, so we were completely dependent on their oral descriptions to reconstruct where they were and what they were doing. Our simulations at Flagstaff and at other locations once again paid off, and we produced a map of the landing site in the SSR based on their descriptions and dead reckoning of how far they traveled between sampling stations.

In addition to the sample collecting, a major objective of Apollo 12 was to land near enough to Surveyor 3 to allow the crew to walk to it and take pictures of the landing site for comparison with the Surveyor TV camera pictures sent back to Earth two and a half years earlier. They would try to bring back pieces of the spacecraft, including the TV camera mirror and scoop, so we could study the effects of thirty months of exposure to the lunar environment. The trajec­tory engineers in mission control and Pete’s piloting skills put the LM right on target, within a few hundred feet of Surveyor 3. This demonstration of the ability to land at a precise point on the Moon, as opposed to Apollo 11’s overshooting the landing point, eased some of management’s concerns as we advocated more difficult future sites. All objectives of the mission were met, and the ALSEP became the first link in the network that the geophysicists had dreamed of for over five years. By the end of their two EVAs, Conrad and Bean had successfully deployed the ALSEP (they encountered a minor difficulty while removing the fuel cask of the radioisotope thermoelectric generator from its stowage on the LM, but deployment proceeded as planned), retrieved pieces from Surveyor 3, and collected a wide variety of samples totaling some seventy – five pounds.

While Conrad and Bean were on the lunar surface, Dick Gordon, the CM pilot, was carrying out his tasks. Soon after the others landed he used his sextant to search for the LM on the surface and was successful, even observing the much smaller Surveyor 3 a short distance away. His primary job was to photograph the Moon from orbit using a Hasselblad and a new camera array called the Multispectral Photography Experiment. The array consisted of four 70 mm Hasselblad cameras with fixed focus, each equipped with a different filter to return photographs in the blue, red, green, and infrared portions of the optical spectrum. This camera array was flown originally on Apollo 9 with Paul Lowman as PI. (For Apollo 12, Alex Goetz of Bellcomm was PI.) Gordon would point the array through one of the CM windows and trigger all four cameras simultaneously every twenty seconds. The major objective was to photograph potential landing sites and, we hoped, use the pictures to extrapolate the re­turned samples to wide areas of the Moon based on spectral differences caused by compositional variation in the lunar soil and rocks. A good concept, but the Moon was not cooperative. When the photographs were developed subtle dif­ferences between the crater Lalande and Mare Nubium were found at only two points. We would have to wait until the J missions, with their more sophisti­cated sensors, to have this exploration technique pay off.

During debriefings of the Apollo 12 crew we asked why they had moved some of the rocks they sampled before documenting their location with photo­graphs, the preferred technique. Their answer was simple and logical. During their early sampling, they had found that many of the rocks they picked up and had documented were too large to fit into the sample bags. Because they were half buried their full size could not be estimated—they were like ‘‘the tip of an iceberg.’’ Rather than waste time photographing samples they could not save, they elected to lift some of the rocks before taking the requested six photo­graphs. As a result of this crew observation, the photo documentation require­ment for the next mission, Apollo 13, was reduced to five per documented sample (although that crew never had the opportunity to use the new standard) and continued to be revised, downward for subsequent missions as we better understood the documentation needs for mapping and cataloging the samples in the LRL.10

Although Gene Shoemaker was still officially the PI for field geology on this mission, Gordon Swann took over crew training and led the interaction of the Field Geology Team with the crew. (Swann would later be named PI for Apollo 14 and Apollo 15.) We exercised the crew at the Cinder Lake Crater Field simulation site outside Flagstaff, described in chapter 9, and other sites, and by mission time Swann and his team had established a good relationship with Pete and Al. They had both been good students, and their training carried over to the lunar surface. In addition to Pete’s enthusiastic, nontechnical descriptions of what he saw, he and Al also provided a good specific commentary that we easily followed, and the Field Geology Team was able to construct a real-time geologic map of the landing site.

After the mission returned we received a letter from a research physicist at the Atomic Energy Commission’s Lawrence Radiation Laboratory in California highly critical of the astronauts’ oral descriptions and their apparently poor training. We always responded to letters from the public on any subject. I was assigned to write a letter back for Rocco’s signature, and it seemed clear to me that the criticism was based on the press reports of Conrad’s voice transmis­sions, not on the whole transcript.11 In the response I included some of the astronauts’ descriptions not carried by the press, such as the characterization as “granitelike” the various colors they reported, and many other precise descrip­tions of rock shapes and soil conditions on the lunar surface. I hoped our response was reassuring to this concerned taxpayer. It was meant not to belittle his concerns but to show that this aspect of the missions—the astronauts’ geological training—was being seriously pursued so that based on their obser­vations we could extract a vast amount of information from each mission.

Apollo 12 had already gone to the Moon and returned before we were pre­sented with the detailed analyses of the Apollo 11 lunar samples. This delay was dictated by the quarantine requirements and by an agreement with the sample PIs not to release their findings until a formal conference could be held in January 1970, when all the results would be available.

Two months later, in March 1970, a new solicitation was issued that required scientists wishing to analyze lunar samples to submit, or resubmit, proposals to receive samples returned by Apollo 14 and subsequent missions. John Pomeroy joined our office at this time to manage the expanded sample analysis program and oversee the operation of the Lunar Receiving Laboratory. By July we had received 383 proposals, including proposals from 175 of the 193 teams (the number had grown from 142) that had analyzed samples from Apollo 11 and Apollo 12. Foreign interest in doing analyses was also growing, and of the 208 new proposals, 95 were from foreign investigators. Gerald ‘‘Jerry’’ Wasserburg, a sample PI from Caltech, writing to administrator Tom Paine in June about his recent trip to Europe, reported that ‘‘there is a fantastic amount of enthusiasm by all the scientists who are involved in these different countries, and. . . the foreign press has given them a tremendous amount of coverage. Some individ­uals, in fact, have become sort of national heroes.’’12 As before, almost all the proposals received were accepted, and many of these investigators and their successors still attend the annual conferences at the Lunar and Planetary In­stitute in Houston.

Before any of the missions, toward the end of 1964 I proposed to NASA management that we study the possibility of commanding the discarded LEM ascent stage to strike the Moon near seismometers that would be placed on the lunar surface by future astronauts.13 At the time, there was no plan to control the impact point of the ascent stage; if not controlled, it would gradually lose altitude and hit the Moon at some unknown time and place. If we could control the impacts of the LEMs, we would have the equivalent of large explosions that would be recorded by the network of seismometers we hoped would soon be in place. We could not be sure when a moonquake or a meteor might provide an energy source large enough to let us study the Moon’s interior. The seismome­ter packages would have finite lifetimes to record some large natural event; if such events were rare, and if the seismometers malfunctioned, they might not be operating when one occurred. Also, the ascent stage was a rather flimsy, lightweight structure, and I feared its impact might not be recorded if its natu­ral decay from lunar orbit occurred some distance away or perhaps even on the Moon’s farside.

We began to explore this idea with MSC and enlisted the support of Frank Press, Bob Kovach, and Maurice Ewing, all members of the seismic teams. It took several years to obtain approval for this maneuver, but by the time Apollo 12 flew we had an agreement to control the impact point of the ascent stage by using the fuel remaining after rendezvous to make it leave orbit at a planned point. For Apollo 12 we recorded the astronauts’ movements and LM takeoff on the ALSEP seismometer as we had for Apollo 11, after which the Moon settled down again and was quiet until the ascent stage hit five hours later, about forty miles away.

We calculated that the impact was equal to setting off an explosive charge with an energy equivalent of about one ton of TNT. The first seismic wave arrived at the Apollo 12 ALSEP 23.5 seconds after impact, building to a maxi­mum amplitude about seven minutes later, with the total recorded event lasting some fifty minutes. The signal recorded was unlike any seismometer recording observed on Earth after either a manmade or a natural event, especially if one considered the relatively small amount of energy involved. This led to a number of theories about the unusual composition of the Moon’s outer layers that might cause such a response. We would have to wait for more information, gathered by the next ALSEPs, before a model of the Moon’s interior finally emerged that most geophysicists would agree with. When we described to George Mueller the effect of the LM impact and the unusual response, he said, tongue in cheek, that the large amount of titanium found in the lunar samples suggested the Moon must be a hollow titanium shell—a spacecraft from an­other galaxy covered with cosmic flotsam and jetsam.

At the end of 1969 Mueller resigned. He had steered the Office of Manned Space Flight, and NASA, to its improbable goal of landing men on the Moon and bringing them safely back to Earth. His management skills have been described by many, and I hope I have given a few insights that will add to an appreciation of those skills. Like Rocco Petrone, he embraced the importance of ensuring that good science be accomplished on the missions. Although I have never been able to ask him why he left NASA, I would not be surprised if a major reason was his frustration at failing to persuade the political powers to approve a long-range plan for continuing manned exploration to the Moon and Mars using the capabilities he and many others had worked so hard to build.

He was replaced by Dale Myers, who had been North American Rockwell’s manager for its Apollo spacecraft contract. Dale had survived both the bad times at Rockwell, when the contract was in trouble for many reasons, and the good times starting with the success of Apollo 8. It must have been a major culture shock to move from being a contractor who had to bow to his NASA ‘‘bosses’’ to being in charge. But he handled it well, and he had a seasoned team to lean on in his first days. I participated in a number of briefings for him early on, and we hardly skipped a beat as we brought him up to speed on all aspects of the program. He selected Charles Mathews as his principal deputy and Charles Donlan as his technical deputy; both were old NASA hands who could help him understand some of the pitfalls he faced. Eight years later, after we had both left NASA, our professional paths would cross again when Dale was appointed undersecretary of the newly created Department of Energy and I was his acting assistant secretary for conservation and solar energy.

Myers’s appointment was only one of several major senior management changes made at this time. Other new blood included George Low, whom Tom Paine, Webb’s successor, brought to Washington from MSC to be deputy ad­ministrator. All these changes had little effect on the upcoming flights. It did seem, however, that once in Washington Low became more sympathetic to the needs of the scientific community, and he strongly supported the efforts to place a high priority on the scientific returns from the final missions.

Once we had an agreement to control the impact of the LM ascent stage, after Apollo 11’s successful mission I proposed deliberately targeting the upper stage (the SIVB) for a lunar impact. This was a lot harder sell than controlling the impact of the LM ascent stage. The SIVB stage, as described in chapter 5, was programmed to deliberately miss the Moon. If it was maneuvered for an impact after placing the CSM and LM on a translunar coasting trajectory, it would arrive at the Moon about the same time the astronauts would be braking into lunar orbit. This was why the original mission rules called for the SIVB to be steered away from the Moon after translunar injection, to avoid any chance that it might interfere with the CSM and LM.

Asking that these rules be changed raised several safety concerns. Not only would the CSM with attached LM and the SIVB be traveling near each other toward the Moon, but it was feared that the powerful impact of the SIVB might hurl debris high above the Moon into the path of the CSM and LM. We asked MSFC to determine if sufficient propulsion would remain after translunar injection so that we could steer the stage and if there would be any problems sending commands to control its trajectory. Douglas Aircraft Company, the SIVB manufacturer, had studied such an application of the Surveyor translunar insertion stage when it was thought that the Surveyor spacecraft would carry seismometers to the Moon, so some of the homework had already been done.

MSFC came back quickly with an analysis that it could be accomplished; it was only too glad to have this opportunity to demonstrate its engineering prowess and the versatility of one of its babies. At the end of May 1969 MSFC made a presentation to me and Michael Yates, and at the end of June we presented our case to the Change Control Board, providing the analyses show­ing that the SIVB could easily be commanded to hit at a preselected point and that debris from the impact would not threaten the LM and CSM.14 Approval was given to proceed with the SIVB modifications, to the delight of the passive seismometer team. We would have to wait until Apollo 13, scheduled for an

April 1970 launch, before all the changes could be made to the SIVB and its command software to achieve the controlled impact.

After Apollo 12, the ‘‘rump GLEP’’ and GLEP came into conflict with the conservative MSC engineers. Some of the sites on our list for the remaining eight missions would require maximum performance from all the Apollo com­ponents. I can recall a contentious meeting at MSC, shortly after Apollo 12’s return, when the subject of future landing sites was on the agenda. This was a meeting of MSC managers and engineers to which a few of us from headquar­ters and Bellcomm were invited. Bob Gilruth, MSC center director, was the senior manager present, but the meeting was run, as usual, by Chris Kraft, Gilruth’s newly appointed deputy, and by Jim McDivitt, manager of the Apollo Spacecraft Program Office. Jim, a recently retired astronaut, was an excellent manager and ran a tight ship. Among other qualities, he was noted for his famous (or infamous, depending on your point of view) daily status reviews, held in a conference room lined with displays and charts and devoid of chairs: no nonsense, get the information out, assign actions, and get back to work! The only bow to comfort was a long table down the middle of the room where you could set your coffee cup while you took notes. Based on his positive response to Petrone’s earlier letter, we considered Jim relatively neutral in our debates on how to accomplish the best science. His major concern was always crew safety; if safety was not compromised, he would usually support our requests.

After the near pinpoint landing of Apollo 12, some of the constraints on site selection described in chapter 5 were relaxed, in particular the requirement for multiple sites to accommodate possible launch aborts. Only one backup site had been designated for Apollo 12, about thirteen degrees farther west, which would have allowed for a one-day recycle if a problem had occurred before launch. The rump GLEP and GLEP went through a process similar to our earlier deliberations to select high priority sites for landings after a successful Apollo 12. This time we came up with a new set A including seventy-two sites. We then narrowed the list to a set B of twenty-one sites and finally recom­mended twelve that included Fra Mauro for Apollo 13 and even more challeng­ing sites for missions 14 through 19. (By now the number of landing missions had been reduced by one, but we were still planning on a total of nine landings.)

But back to the meeting. Equatorial sites had been agreed on for the first three landings as the safest and most easily accessible, although the Apollo 13 site, Fra Mauro, would be a little more challenging since it was surrounded by

rougher terrain. These initial sites were within the “Apollo zone of interest.’’ All were close to the equator and were covered by the greatest number of high resolution Lunar Orbiter photographs. Many uncertainties still existed in pre­dicting the performance of the total Apollo system, but Bellcomm had already completed an analysis of SIVB, LM, and CSM performance showing that a high percentage of the Moon’s nearside could be reached while maintaining the required safety margins.

As the meeting droned on and such things as communication restrictions and propulsion budgets and margins were discussed, it became apparent that MSC management was going to take a conservative stand. Those of us who had been working on future landing sites were being asked (not quite directed) to rein in our expectations and continue to look for science sites near the Moon’s equator. The nearer to the equator you landed, the more options were available to get you out of trouble. There was reluctance to go outside the ‘‘Apollo zone’’ despite the Bellcomm study. Besides, it was a Bellcomm analysis, not one done by MSC engineers.

MSC’s position was certainly understandable. Every mission was risky, from liftoff to splashdown, and a difficult lunar landing site only added to the risk. No one wanted to be responsible for the decision to land at a site where a crew would be lost, for whatever reason. An accident, such as befell the crew of Apollo 1, could result in the cancellation of the remaining missions, an outcome that few in NASA would have cheered. For the staff at MSC each flight involved more personal worries than, perhaps, for someone in Washington or elsewhere in the scientific community; crew members were their neighbors and co­workers. If a crew didn’t return they would be living with the grieving families.

By this time I had many close friends in the astronaut corps and fully appreciated the danger inherent in each mission. However, Noel Hinners and I felt obliged to speak up. The only rationale for continuing the missions was to carry out good science, and this could be done only if we were allowed to explore sites far from the equator, sites already identified as having the potential to resolve important questions. We went so far as to predict that, based on Lunar Orbiter photographs, safe LM landing sites could be found almost any­where on the Moon. If any other proof was needed, look at Surveyor 7, which, with minimum ability to target the landing site, had managed to land in rough terrain on Tycho’s rim without an astronaut making last-minute adjustments. How much easier it should be with a man at the controls. These remarks were met with skepticism and grumbling from around the table, but this position was gaining support from many others, including some of the astronauts.

Eventually, as others with more clout weighed in, MSC management reluc­tantly agreed to process sites away from the equator. Undoubtedly each mission that lifted off after Apollo 13’s near disaster increased their anxiety; the chances of a major problem were rising with each flight. No matter how carefully we prepared, one or more of the five million parts included in every launch vehicle and spacecraft could fail or malfunction at any point in a mission.

On March 6, 1970, the Apollo Site Selection Board met at KSC to select the landing site for Apollo 14. With Apollo 13 scheduled to land in the western part of the ‘‘Apollo zone,’’ this was the first meeting of the board since the meeting described above. We looked on it as a test to see if MSC management would be swayed by our arguments and allow Apollo 14 to land outside the ‘‘zone.’’ Tony Calio, who had replaced Bill Hess as chairman of GLEP, presented the results of the GLEP meeting of February 6 and 7. GLEP recommended a site called Littrow, at the southeastern edge of Mare Serenitatis, well north of the ‘‘Apollo zone,’’ and the MSC in-house site evaluation team recommended the same site. After several presentations, including two by Lee Scherer and Noel Hinners, the board approved the Littrow landing site, and Jim McDivitt signed off in agree­ment.15 We had overcome the last hurdle toward planning the scientific explo­ration of the Moon during Apollo.

A key science ally at MSC was Jack Sevier. Jack’s personality was perfect for the difficult job he was assigned, acting as a mediator between the scientists and MSC’s engineers. Easygoing, with a ready smile and quiet sense of humor, Jack had been an important contributor to the rump GLEP meetings starting in 1967, providing MSC’s views on the constraints that could affect site selection. He was the branch chief of the Operations Analysis Branch and as such was the focal point for all the competing factors that could influence the outcome of our scientific activities. He would later lead the Lunar Surface Planning Team for the J missions, which developed the astronauts’ lunar surface timelines and ultimately shaped the successful outcome of each EVA.

With the missions still being scheduled rather rapidly and changes in their scientific content occurring with each mission, some members of the scientific community continued to publicly criticize how Apollo science was progressing. Soon after the return of Apollo 11, Gene Shoemaker was quoted as being highly critical of the way NASA management treated science on the Apollo missions.

This view troubled me deeply at the time: we had been working hard to expand the science, and I knew he was aware of how much more productive the next missions would be. There is no question they could have been better, but we had made great progress since he had first become involved. His statement drew the ire of Homer Newell and Rocco Petrone. Harold Urey, perhaps egged on by Tommy Gold, who always seemed to delight in knocking NASA, also criticized the lack of scientific input into NASA decision making.

In a letter to Newell in March 1970, Urey said he agreed with Gold ‘‘that well known people who have been concerned about the moon for years are so systematically neglected by the management of NASA.’’16 He was particularly irate at their exclusion from the selection of Apollo landing sites. In regard to site selection he wrote that ‘‘the people who vote are loaded with geologists of a very limited view of lunar science,’’ and he made a few other scathing com­ments. By this time, after just two missions, Urey was seeing the writing on the wall. His well-publicized theories on the Moon’s origin were being proved wrong, and I suppose Nobel laureates don’t like to be proved wrong. At a later date he would acknowledge his errors and even make jokes about them.

Newell’s staff was asked to respond to Urey’s letter, but they sent an informa­tion copy to our office. Rocco Petrone, not taking kindly to this criticism, asked that we address one of Urey’s comments dealing with site selection. In my memo for Petrone’s signature, which we hoped would be included in Newell’s formal response, I named the scientists and engineers present and voting at the last Site Selection Subcommittee meeting at MSC.17 I listed twenty-one names: three geologists, two astronomers, four geophysicists, three NASA engineers, two geochemists, one nuclear chemist, three physicists, one geodesist, one atmospheric physicist, and one cosmologist-chemist, Harold Urey. Of the twenty-one, nine were government employees or contractors and the other twelve came from universities or private research laboratories. All had been involved in lunar research for at least the past five to ten years, which pretty well covered the period when interest in the Moon became widespread. Urey had picked the wrong topic—site selection—to complain about, but his overall concern had some merit. His complaints and those of others were primarily a criticism of how MSC was interacting with the scientific community, which once again was becoming contentious after Tony Calio replaced Bill Hess.

Urey’s letter came just one month after a meeting at MSC when a group of

scientists, all closely involved in Apollo investigations, met with MSC manage­ment to discuss the problems they were having working with MSC staff. Urey had not been invited to this meeting, nor had Tommy Gold, which may have added to their pique; Newell attended as an observer. After the meeting Newell apparently thought the situation was resolved and wrote Gilruth a complimen­tary letter; but he didn’t really understand the depth of distrust that was build­ing between Calio’s organization and the scientists who were devoting more and more of their time to making each mission as successful as possible. Yet the meeting was useful in making McDivitt and Chris Kraft more aware of the needs of the scientists, and relationships with their offices improved. MSC agreed to arrange for more time in the astronauts’ schedules so the PIs could explain their experiments and their requirements during deployment or opera­tion. The PIs also asked for a better system of communication between the scientists in the SSR and the crews. They cited difficulties that arose during Apollo 12, when it took ten to fifteen minutes for questions raised in the SSR to be relayed to the astronauts by the CapCom, if they went out at all—and often they didn’t.18 There was some improvement on succeeding missions, but in general MSC and the Flight Operations Directorate (FOD) tended to ignore this latter request. Flight directors and CapComs felt, with some justification, that they shouldn’t interrupt the crews on the surface with a lot of questions and directions; they had enough to think about.

Apollo 13 was a scientific disappointment but an engineering triumph. We lost a precious ALSEP (one of only six purchased), but the opportunity to study this site and collect valuable samples was realized when Apollo 14 went back to the Apollo 13 landing site. In spite of this disappointment, I never heard any complaints from the PIs, many of whom had worked with Jim Lovell, Fred Haise, and John ‘‘Jack’’ Swigert to prepare them for their flight. Like everyone else, we could only cheer the skill of all the NASA engineers and support contractors who brought the crew home safely. The Apollo 13 crew members who performed so well under the threat of being the first astronauts to die somewhere in space, and the many heroes in the FOD led by Eugene Kranz, have had their roles well documented, so I will not try to add to that story. Science probably gained from the failed landing. It helped us refocus on how important each mission was. There were no givens; we had to make sure the remaining missions would be fruitful. And it seemed to make management more receptive to our requests to improve the science content of the last mis­sions. The drama of Apollo 13’s rescue also ensured a more attentive public for the next missions and a wider audience interested in what we were discovering.

One experiment, the passive seismometer left behind at the Apollo 12 land­ing site, did achieve important results from Apollo 13. Despite the problems the crew encountered during the rest of the mission, the Apollo 13 SIVB stage, the first programmed to strike the Moon, accomplished its job by landing some eighty-five miles from the Apollo 12 ALSEP. The seismometer received strong signals, and the impact had so much energy-estimated to be the equivalent of twelve tons of TNT (larger than the LM ascent stage impact because of its greater mass and higher velocity at impact)-that it sent seismic waves deep into the lunar crust. This elated Gary Latham, the passive seismometer PI, because he and his coinvestigators could now make some preliminary estimates about the Moon’s deep structure.

When Lee Scherer’s office was formed at the end of 1967, several of us involved in lunar science planning left Advanced Manned Missions, but Phil Culbertson stayed, eventually becoming director. In March 1970 he negotiated a memorandum of understanding with the Apollo Program Office to work cooperatively on lunar planning in case funding became available to continue missions beyond the scheduled Apollo flights.19 Our two offices continued working jointly on post-Apollo planning for several more years, despite the lack of official sanctions to build the hardware needed for extended missions.

After Apollo 13 failed to land, and reacting to the increasing clamor in some circles to halt the missions, in July 1970 our office issued a summary report of what we had learned to date from all our missions, manned and unmanned, and where we thought lunar exploration should be going.20 The objective of the report was to support Culbertson’s planning efforts in Advanced Manned Mis­sions and to present an “Integrated Space Program Plan’’ that would provide mission schedules extending to 1990. It represented our last effort to justify a continuing program of manned and unmanned exploration by building on Apollo and other programs, including Mariner and Viking, and factoring in programs on the drawing boards such as Skylab and space stations. We pre­sented an integrated program that included lunar bases and manned inter­planetary launches.

Recently a quotation from Charles Lindbergh came to my attention. Asked about the $25,000 Orteig Prize offered for the first nonstop flight between New

York and Paris, which he won with his daring flight in 1927, he responded, ‘‘The important thing is to start: to lay a plan, and then follow it step by step, no matter how small or large each one by itself may seem.’’ One could make a reasonable argument that Lindbergh’s successful flight was the first step toward today’s commonplace travel across the Atlantic and to almost every point on the globe. With Project Apollo we had taken the first step in mankind’s leaving Earth and exploring our solar system. We believed we had put forth a step-by­step program to build on Apollo and move logically to the next objectives: space stations, lunar bases, and manned flights to Mars as early as 1989.

No such logical plan was ever agreed to. Some administrations have ignored space exploration, and some have paid it lip service. In the end, a program that would take advantage of the expertise and capabilities developed for Apollo was never endorsed. The report is now resting in one of my dilapidated packing boxes, perhaps the only surviving copy of our vision of a long-range plan for exploring the solar system. It was grandiose—undoubtedly too grandiose for the times—but in 1970 everything we proposed was achievable based on the technology in hand. All that was needed was the leadership to commit the nation to the next step.

In January 1971, just two weeks before the scheduled liftoff of Apollo 14, the second lunar science conference was held at the Lunar Science Institute. Al­though many of the same people attended this conference as were at the one held after the study of the Apollo 11 samples, the sense of excitement was missing. The only new samples that had been studied, aside from a few grams of material brought back by the Soviets’ Luna 16, were those returned by Apollo 12 a year earlier. Whereas restrictions had been placed on the release of infor­mation about the Apollo 11 samples, the Apollo 12 sample PIs were not pre­vented from publishing the results of their studies of material returned by the mission. Most of the new information was already public and well known by the attendees.

The big debate at the conference dealt with the significance of the high content of radioactive elements (uranium, thorium, and potassium 40) found in some of the Apollo 12 samples, which would imply an early, very ‘‘volcanic’’ Moon. There were other differences from the Apollo 11 samples, suggesting that the Moon may have had an unusual differentiation history. It also appeared after initial study that the mare material sampled at the Apollo 12 site was about a billion years younger than that collected at the Apollo 11 landing site, suggest­ing that the Moon had gone through several major periods of mare formation. These findings would continue to be debated as each mission brought back new information.

Apollo 14 was sent to the site chosen for Apollo 13, Fra Mauro, in a hilly, upland area just a short distance east (112 miles) of the Apollo 12 site. From the perspective of our plans to deploy the ALSEPs in a broad network so we could triangulate on phenomena at the Moon’s surface or occurring at depth, being so near the Apollo 12 ALSEP was not ideal. But from a geological point of view it was considered an important site, since we believed that the samples returned would include debris ejected from the huge Imbrium basin to the north. Again, as for Apollo 12, we hoped to collect samples from deep within the Moon that would help resolve some of the questions raised at the second lunar science conference. They would also be useful to Gary Latham and his coinvestigators in interpreting the Moon’s deep structure, since these rocks would tell them how fast the seismic waves created by the SIVB impacts should travel compared with what they were observing in the records received back on Earth.

Had Apollo 13 been successful, we were willing to accept the deployment of the ALSEP so close to Apollo 12. It was to be the last of the landings near the Moon’s equator, reflecting MSC’s cautious approach. We had expected that after Apollo 13, Apollo 14 would land at Littrow, the first site selected solely for its scientific value and, because it was far off the lunar equator, ideal for our ALSEP network. The geological rationale for landing at Fra Mauro still held, but the decision to retarget Apollo 14 there was doubly painful from a scientific perspective. With the loss of Apollo 13, there were only six more projected landings (ultimately reduced to three) to uncover the Moon’s secrets hidden on or below a surface area roughly equivalent to all of North and South America combined. And well over half of that area was inaccessible because it was outside our landing capabilities or on the Moon’s farside. Imagine trying to understand those two continents with only six widely scattered small points of knowledge!

The landing site was to be within walking distance of what appeared to be a crater of recent vintage, named Cone by the Field Geology Team because of its steep, funnel-like inner slopes. From the Lunar Orbiter photos we could see large blocks on the rim of Cone Crater, reinforcing the belief that if the astro­nauts could get to the rim they would be able to sample Imbrium ejecta in the rocks ‘‘mined’’ by the Cone Crater impact. Alan Shepard guided the LM to a perfect landing within two hundred feet of the target point and less than a mile from Cone Crater, whose rim could be seen in the distance when he and Edgar Mitchell descended the LM’s ladder. This time there was a color television camera, with better resolution than the Apollo 11 camera, and it functioned well, providing views of the astronauts as they climbed down to the surface and panoramas of the landing site as they worked near the LM.

Between the Apollo 13 and Apollo 14 launches we had built a small two­wheeled cart, the modularized equipment transporter (MET) discussed in chapter 8, to help the astronauts carry all the gear that was now part of the field geology experiment. It was unloaded from the LM descent stage near the begin­ning of the first EVA, and the crew stowed the tools and equipment they would need for the sampling scheduled on the first EVA and the traverse to the rim of Cone Crater, the major objective of the second EVA.

The first EVA went off with no big hitches, and the major tasks—the ALSEP deployment and sample collection near the landing site—were successfully completed. A new experiment, the active seismic experiment, was conducted in conjunction with the ALSEP deployment. Three geophones were strung out on cables to the south of the ALSEP, with the last one approximately three hundred feet from the ALSEP central station. The first part of the experiment consisted of setting off small charges, about the size of a shotgun shell, housed in a hand­held ‘‘thumper’’ hardwired to the ALSEP central station electronics, which provided timing data and transmitted the signals received by the geophones back to Earth. Mitchell carried the thumper out to the last geophone and set off a charge, then retraced his steps back to the geophone closest to the central station, setting off charges along the way. Twenty-one charges were scheduled, but a few misfired and only thirteen were recorded. A second part of the experiment consisted of a mortar designed to fire four small explosive charges various distances away from the geophones, the farthest to land five thousand feet from the mortar. This second part of the experiment was not conducted until many months later, to avoid any possibility that the mortar fire might damage the nearby ALSEP central station.

Although Shepard and Mitchell could see the ridge formed by Cone Crater in the distance when they started out on the second EVA, once they began walking and pulling the MET they soon lost sight of the ridge behind the intervening low hills and hummocks. Others have described in some detail their difficulties in reaching the rim of Cone Crater. They didn’t quite make it, but they came close, and they sampled ejecta thrown out by the impact that formed the crater, the main geological objective of the mission. After the diffi­culties they encountered attempting to reach Cone Crater’s rim, they probably both wished they had the LRV that would be carried on the next mission. Another new experiment on this mission, the Lunar Portable Magnetometer, was operated twice during this EVA, and the readings were relayed back to Houston by voice. The samples collected during both EVAs weighed almost ninety-five pounds.

Like the CM pilots before him, Stuart Roosa carried out several experiments on the way to the Moon and while the other astronauts were on the lunar surface. The number of experiments assigned to the CM pilot was increasing with each mission as we attempted to take full advantage of his time and the added payload weight that was becoming available. Roosa completed several new photographic tasks and other types of experiments. Bellcommers Farouk El Baz and Jim Head took on growing roles instructing Roosa, as well as the CM pilots on the final three flights, in the objectives of the photographic experi­ments and the cameras’ operation. After the film was returned, they also helped interpret the data obtained. Apollo 14 marked the end of the H missions, one short of the four originally planned.

The J Missions: We Almost Achieve Our Early Dreams

Apollo 15 was the first of the J missions. Years of struggle and cajoling, as well as long hours spent meeting with contractors, principal investigators, scientific committees, and NASA colleagues, had finally borne fruit. All the allowances for payload, extravehicular activity time, distance traversed, and sample return would suddenly double or triple. Although we had greatly increased our ability to explore, however, there would be no dual launches, no two-week exploration timelines to construct, and no seven-thousand-pound science and logistics payloads that would have given us the experience to plan for lunar bases. Our attempts to convince Congress and the Nixon administration to extend lunar exploration had failed. Instead of the five more missions we had been planning just six months earlier, only the three J missions remained. After Apollo 17 returned, Project Apollo would close its doors. We chose to put this sad ending out of our minds and concentrate on ensuring the success of the last missions.

Two days before the launch of Apollo 15 several of my colleagues and I flew to Orlando and then drove to Cocoa Beach, Florida, to prepare for the prelaunch press briefing. George Esenwein and Floyd Roberson would describe the new command and service module experiments; Ben Milwitsky and Richard Diller would do the same for the lunar roving vehicle; and I would cover the surface science. The routine at these briefings was that we would make short prepared statements, illustrated with vugraphs, and then take questions. Gene Simmons joined us, having recently transferred from MIT to the Manned Spacecraft Center, and some of the PIs—those with experiments flying for the first time— were on hand to discuss them. Gene, with his new title of chief scientist and an assignment to once again try to improve relations with the scientific commu­nity, had written a guidebook, ‘‘On the Moon with Apollo 15.” It was sought after by the media as a quick reference covering the science aspects of the mission and providing some easy quotations, a service the hardworking mem­bers of the press always appreciated. Gene compiled similar guidebooks for the last two missions, Apollo 16 and Apollo 17.

NASA’s Public Affairs Office usually released a mission press kit about ten days before the launch to give the media a chance to get familiar with the mission. It included details on all aspects, including the mission’s scientific activities and experiments. Before these last briefings at Cocoa Beach, media briefings for each launch were conducted at intervals at places such as MSC and Kennedy Space Center, and a major briefing was always held at headquarters about a month before launch for the large Washington media contingent. But the various briefings at Cocoa Beach (other parts of the flight besides the science were covered) the day before the launch were always the best attended, resulting in a lot of print and sound-bite coverage. Some seventy-five members of the media attended our briefing, held in a large second-floor conference room at the Friendship Inn the morning of July 25. Jack Hanley and Don Senich came along to answer any questions on two new pieces of equipment, the lunar drill and the soil mechanics penetrometer.

By this time our office staff had grown, mostly with members detailed from other agencies or NASA centers. NASA budgets had been going down for the past four years, and with those reductions came a semifreeze on hiring NASA civil servants. We always argued, to no avail, that some aspects of NASA busi­ness were still growing—Apollo science as an example—and needed more bodies. To spread the added workload in our office, we obtained detailees from the Army Corps of Engineers, the United States Geological Survey, and the Jet Propulsion Laboratory. In addition to Jack Hanley from USGS and Don Senich from the Corps of Engineers, USGS lent us Gerald ‘‘Jerry’’ Goldberg, and JPL sent Ewald Herr and later Peter Mason and Ronald Toms for assignments that lasted one year or more. Hanley, Goldberg, and Senich stayed with our office for over three years and were invaluable additions, cheerfully (usually) taking on the ‘‘dog work’’ that every government bureaucracy generates as well as the more interesting oversight for science payload development.

Besides the excitement of the upcoming launch, which brought media rep­resentatives from all over the world, in the days before the launch Cocoa Beach and the surrounding area were the site of many parties, a tradition that went back to the first rocket launches in the late fifties. These parties grew with each new manned launch. By custom, the night before the launch the big companies, with a major stake in the mission, would hold open houses that included food and drink. For Apollo launches, North American, Douglas Aircraft, Grumman, and Boeing, as well as smaller companies such as Bendix (many of these com­panies have since merged and lost their identities), would all hold their own affairs to tout their participation, with some competition to throw the best party. Similar “splashdown” parties were held in Houston, near MSC, after the end of each mission, and these would be even wilder, if that was possible, than the prelaunch parties at the Cape. These MSC parties were by invitation only, and invitations were always in great demand.

Up and down the beach a half dozen or more parties would go on into the wee hours. The morning after would be spent describing some of the more outrageous events. Nothing attracts the press more than a free party and, I might add, the many VIPs and sightseers who were in attendance. We civil servants tried to be discreet, but we would also drop in on a few of the parties even though some of us had official duties the next morning. With the Apollo 15 launch scheduled for 8:34 a. m., it meant waking up early to beat the traffic and get to my VOA broadcast site. I had been promoted since Apollo 11 and was now a full partner in the broadcasts.

The TV networks were getting more and more elaborate with their coverage of each succeeding mission. For Apollo 15, with its promise of real-time TV pictures during the astronauts’ LRV traverses, the major networks had assem­bled simulated lunar terrain to illustrate how the astronauts were going about their exploration. NBC had a small working model of the LRV, and CBS and ABC had borrowed full-scale working models. For TV and news media pools not fortunate enough to have their own models, we supplied static mock-ups of the LRV and Apollo Lunar Surface Experiments Package at MSC, where their reporters could be shown standing in front of the models.1 Each of the major networks also had a captive astronaut in the studio to explain the intimate details of what was going on during the mission.

In terms of science training, the crew of Apollo 15 was the best prepared yet. Based on my observations, Dave Scott, James Irwin, and Alfred Worden showed the greatest interest of any of the crews to date in understanding the science objectives for their landing site—the Apennine Mountains and nearby Hadley Rille—and the new suite of experiments housed in the lunar module and CSM.

Scott, as mission commander, set an example for his two crewmates through his hard work and contagious enthusiasm. Hadley Rille, a long, sinuous valley, was one of the most intriguing features on the Moon’s surface. It was surpassed in interest only by the first landing site, when any information returned was bound to be extraordinary. Several theories had been proposed to explain the rill’s origin: that it had been formed by water discharged from the Moon’s interior (the least favorite theory among most lunar scholars); that it was a lava channel or collapsed lava tube or had been gouged from the surface by some volcanic event; or that it was the remnant of faulting or stretching of the lunar crust.

The landing site was on the eastern rim of the Imbrium basin, so we antici­pated that the returned samples would include material from the Apennine Mountains, probably formed by uplift and ejecta from deep in the Moon’s interior. Other samples should include Imbrium basin fill consisting of some type of lava. Samples from the edge of Hadley Rille might resolve the question of its origin. We might even collect samples of the ejecta from Mare Serenitatis, just a short distance to the east. If we could identify their source, age dating these samples would go a long way toward explaining key events in the Moon’s early history that shaped its final form.

Adding to our excitement about this mission was the greatly increased radius of operation for the astronauts and the many new experiments that would be performed. The science payload delivered to the lunar surface would be almost 1,200 pounds, compared with the Apollo 11 payload of less than 200 pounds and more recently the 470 pounds carried on Apollo 14. If the mission went as planned, Apollo 15 would come closest to our earlier dreams for the first post – Apollo missions. Apollo 15 marked another milestone: Jack Schmitt was named to the backup crew. Based on previous crew rotations, this would have put him in position to be named to the prime crew for Apollo 18, now canceled, but at least he had moved up in the pecking order. This would be Gordon Swann’s last mission as PI for the field geology experiment. He and his many Flagstaff colleagues and coinvestigators, which included Bill Muehlberger, the designated PI for the last two missions, had been building to this climax after many years of hard work. Next best to being on the Moon themselves, the J missions would validate their efforts, and they worked tirelessly to prepare the crew.

All the preparation paid off. The crew performed flawlessly. The Science Support Room (SSR) geology team, led by Swann, listened, recorded, debated, and attempted to interpret in real time everything that was happening on each of the three EVAs during the 67 hours the crew was on the surface. Total time of the three EVAs was 18.5 hours, a new record. As we had practiced at Flagstaff, Scott also performed the first and only stand-up EVA when, shortly after land­ing, he opened Falcons overhead hatch and stood on top of the ascent stage engine cover to get a bird’s-eye view of the landing site. While enjoying the scene around him, he took some panoramic pictures and planned the upcom­ing traverses. With only minor upgrades of LM systems for the J missions, we were able to more than triple the time the Apollo 11 crew had spent on the lunar surface. The Apollo 15 EVA time was almost as long as the total time the Apollo 11 astronauts spent on the Moon. These numbers confirmed in my mind that the one – to two-week visits we envisioned for the post-Apollo dual-launch missions could have been achieved by making the modifications to the Apollo systems we had studied.

Another change for Apollo 15 was the inclusion of scientist-astronauts Jack Schmitt, Joe Allen, Bob Parker, and Karl Henize as capsule communicators; Allen, who had served as mission scientist, usually manned the console during the EVAs. On all the previous missions only Schmitt, for Apollo 11, had been given this high profile task. Although it is difficult to point to any specific advantages of having them at the consoles, interaction between the ground and the crew of Apollo 15 was lively, and certainly we in the SSR felt more comfort­able knowing that Allen could immediately interact with the crew if necessary. We passed suggestions and questions to the CapComs, and in contrast to Apollo 12, many were passed on.

The scientific harvest from Apollo 15 was spectacular, derived both from the lunar surface and from lunar orbit. An ALSEP was deployed at the north­ernmost point reached by any Apollo mission. This ensured an ideal position­ing of the ALSEPs for triangulating readings for experiments like the passive seismometer and LRRR that needed site separation and the surface magne­tometer that was attempting to discover if the Moon’s magnetic field might vary from site to site. The three LRV traverses covered almost seventeen miles, during which the astronauts studied twelve locations in addition to the imme­diate landing site. They collected almost 170 pounds of samples and took more than 1,100 photographs. Besides the many photographs taken with the Hassel – blad cameras, we also had TV coverage of eight stations and some footage taken while the astronauts were under way on the LRV. Eight major experiments were conducted in lunar orbit, and many types of photographs were taken from both the CM and the scientific instrumentation module (SIM) bay in the service module, adding to the wealth of new information that included coverage of the Moon’s farside.

Apollo 15 recorded one other first, and last. After Scott and Irwin’s ren­dezvous with Worden back in lunar orbit, Lee Silver was called out of the SSR to go to the Mission Operations Control Room. Lee had led many of the field training exercises for the Apollo 15 crew and had established a close rapport with them. Scott wanted to talk directly to Lee to thank him for his long hours and dedication to their education. Thus, for the first and last time during an Apollo mission, a member of one of the science teams talked directly to the astronauts while a mission was in progress without going through an astronaut CapCom. Lee passed on our congratulations and told them how excited we were about what they had accomplished, and he reported the results we were already seeing as we reduced the traverse data. I hope that when we return to the Moon such exchanges between scientists on Earth and those working on the lunar surface will be the norm, for it will surely add to the value and efficiency of future lunar exploration.

All the experiments and equipment carried on Apollo 15 performed up to expectations except for the drill. To quote from the crew’s observations, ‘‘The deep core could not be extracted from the uncooperative soil by normal methods; the two of us, working at the limit of our combined strength, were ultimately required to remove it.’’ The exterior flutes contributed to this condi­tion because the drill stem was pulled into the ground still deeper when the motor was activated.2 This activation was supposed to clear the flutes for easy extraction.

As we were to discover, there were two complications involving the drill. The first and most serious occurred as the astronauts were drilling the bore holes for the heat flow experiment, and they encountered the second while trying to extract the core sample. Mark Langseth, the heat flow PI, had designed his experiment around placing the sensors in a cased drill hole some ten feet deep. When Scott attempted to drill the first hole he could not go much deeper than about five feet, well short of his target. The drill stem refused to go any farther no matter how hard he tried. Frustrated and thinking he might have hit a large rock, he stopped working on the first hole and tried to drill the second. Again, he could not get penetration much below the first length of drill stem. Time was fleeing, so he was instructed to stop drilling, place the sensors in the holes he had, and finish the remaining first EVA tasks. Langseth ended up with his sensors much closer to the lunar surface than he wanted, and he feared he would not get the high quality data he was hoping for. More later on the tribulations of the heat flow experiment.

In drilling the core sample the astronauts encountered a different problem. The drill penetrated to the full depth quite easily—too easily, it turned out. Fortunately, through the crew’s ‘‘combined strength’’ they salvaged the core. Thus the time had not been wasted. Wasting time during a mission was not attractive to anyone, especially when the medical team monitoring the astro­nauts could see they were in danger of exceeding their physical limits while trying to remove the core. As the astronauts struggled, recommendations were made in Mission Control to abandon the attempt, but Scott and Irwin persisted and saved the day.

Immediately after this well-publicized glitch, while the crew was still on the Moon, Rocco Petrone caught me in the Mission Control Center and issued one of his famous edicts—we were to solve the drill problem before the Apollo 16 flight readiness review! I huddled with Jack Hanley to discuss our course of action. Following our usual method of addressing such issues, we appointed a ‘‘tiger team,’’ consisting of Dave Carrier and several other MSC engineers plus Jack Hanley and Don Senich from my office. They were dispatched to Denver to meet with Martin Marietta, find out what went wrong, and make the necessary modifications. The solution had to be found quickly, since the drill had to meet the Apollo 16 equipment stowage window for its preflight checks. We also had to be prepared to modify the astronauts’ training and simulation schedules if any changes required new instructions or training.

After discussing the crew’s observations and reviewing how the drill had performed, we knew we had two separate problems, one in drilling the bore holes for the heat flow experiment and the second in extracting the deep core. The tubular drill sections used for the heat flow holes were of a different design than the core stems. Since the core stems had drilled to almost eight feet without any trouble, this design difference had to hold the clue. Langseth thought he knew what had gone wrong. The flutes on the outside of the heat flow drill sections did not extend the full length of each section; they stopped short of the ends, leaving an open space on the shafts.

With new information on the characteristics of the lunar soil derived from the soil mechanics experiment, the fidelity of the simulated lunar soils used during testing was improved, and we conducted many tests, some in a vacuum chamber. Langseth was right. The short interruption of the flutes at the joints had prevented the cuttings from traveling up the tubular sections; they jammed at the joint after the next section was added and the joint was drilled a short distance below the surface. The design had worked satisfactorily during our terrestrial trials, but the soil simulant used during the original tests had not been as compact and dense as real lunar soil.

The original design of the heat flow tubing had called for titanium inserts at the ends of the fiberglass sections to strengthen the joints so they could be screwed together like the core sections. This would also ensure good meshing of the flutes, but Langseth had been concerned that so much metal in the tubing might disturb the sensitive readings he was hoping to obtain. We had removed the titanium joints, and the tubing carried on Apollo 15 required that each section be pushed into the next. The flute alignment was determined by how carefully the astronauts joined the sections, but there was always a small gap be­tween the flutes. Langseth agreed that we would have to go back to the original design for Apollo 16. This was done, and the drilling tests were successful.

We believed we had solved the heat flow drilling problem, but we needed more tests to understand the core extraction problem. During the crew debrief­ing it emerged that because Scott had had trouble drilling the heat flow bore holes, he had perhaps put too much pressure on the drill while coring. This forced the core stem into the soil before the flutes could completely clear the cuttings, thus jamming the core stem. Tests at Martin Marietta showed that if the drill penetrated more slowly, even in a more moonlike soil than we had used in previous simulations, there should be no difficulty extracting the core on the next missions. Just in case it did jam, we designed a new device to jack the drill stem out of the hole if it would not come free using the normal procedure of rotating the drill core in place. Score two hits for the tiger team and the Martin Marietta and Black and Decker engineers. On Apollo 16 and Apollo 17 the drill worked well for both applications.

Although the missions, starting with Apollo 15, were being launched on a more relaxed schedule, approximately one every eight months, there was a heavy training burden on everyone at MSC, KSC, and the Field Geology Team. At headquarters we also felt the pinch as we tried to stay abreast of the progress, or lack thereof, so we could keep Petrone and other senior management up to date. Rocco hated surprises, and this attitude carried over to all his staff and was often reflected during his weekly status reviews. These reviews, held at our offices at L’Enfant Plaza in a large, windowless room lined on both sides with multiple sliding status boards, would consume half a day or longer depending on the number of outstanding issues. The status boards were updated daily through a contract with the Boeing Company and were used extensively during the reviews. Each Apollo office would make a presentation so that Rocco could get a snapshot of the program covering everything from spacecraft and payload status to funding, manpower, and eventually, final plans for close-out. After the near disaster of Apollo 13 and with the last missions firmly scheduled, the atmosphere was getting tenser; we had to make sure that nothing was left to chance and that no dumb mistake would jeopardize a crew.

In November 1971, four months after the return of Apollo 15, Lee Scherer was transferred to a more prestigious management position, director of the NASA Dryden Flight Center at Edwards Air Force Base in California. Don Wise, Lee’s deputy, tiring of the Washington scene, had already gone back to academia. O. B. O’Bryant was named to replace Lee for the final two missions, and I was named program manager for Apollo surface experiments, including the ALSEP, since Ed Davin and Dick Green had left to take new jobs at the National Science Foundation.

The crew of Apollo 16—John Young, Charley Duke, and Kenneth Mat­tingly—had been named long before Apollo 15 was launched, and their training and simulations overlapped those of the Apollo 15 crew. For the Field Geology Team there was the added consideration of changing the PI from Gordon Swann to Bill Muehlberger, with Bill adding some new coinvestigators from USGS and Bellcomm. This transition went off without a hitch, since both Gordon and Bill had worked together for a long time and there were no profes­sional jealousies involved in the switch. With the overlap in training the Apollo 15 and Apollo 16 crews, it was difficult even to notice a change, and almost all the faces remained the same. With each mission, the training was also becom­ing more complicated, reflecting the added complexities of the J missions with their new experiments, both surface and orbital, and longer surface EVAs.

The Apollo 16 crew had a much different character than the crew of Apollo 15, reflecting the personality of John Young. He had already flown three space missions, including the highly successful Apollo 10, and had more time in space than any other astronaut except Jim Lovell. Young was more relaxed than the hard-driving Scott. He was quick with a quip or story to break the intensity of a training or simulation session. In spite of the more relaxed atmosphere, the crew was required to spend hundreds of hours learning the scientific nuances of their chosen landing site, Descartes, a highlands crater near the Moon’s center several hundred miles southwest of the Apollo 11 landing site. In addition to learning what they should expect and look for at Descartes, they had to train to deploy the ALSEP and all the other experiments, including the redesigned drill and one new experiment, the far UV camera-spectrograph, our first chance to use the Moon for astronomical studies.

Bill Muehlberger, perhaps tempered by dealing with undergraduates during his tenure at the University of Texas, Austin, plus his long association with some of the eccentric personalities at USGS, meshed well with the crew. His team of coinvestigators and field geology instructors took on the task of instructing the crew by taking them to several training sites in the United States, Mexico, and Canada. The sites were chosen to expose them to field conditions representative of the latest geologic interpretations of what they might encounter at Descartes. Since this would be the first landing in the Moon’s highlands, we expected to pin down their composition, an important determination in understanding the Moon’s history. The crew would also sample the material that filled the large crater in which they would land; perhaps it was of a different composition than the maria sampled on the earlier missions. Some photogeologists studying the area around the landing site believed they were observing volcanic features, low hills that might be composed of lava or cinders, much like the formations just east of Flagstaff. Other interpretations were possible, but these hills looked unusual and, most important, were in the highlands. Young and Duke were conscientious students and quick learners. Those of us who tagged along to observe the training sessions were impressed at how well they were absorbing the huge amount of information thrown at them. The sessions included a trip to Sudbury, Canada (which I didn’t attend), considered an excellent example of some of the geological situations they might encounter on the Moon.

Ken Mattingly, the CM pilot, who had been scratched from the Apollo 13 crew because of fears he might have been exposed to German measles, for which he had no immunity, was the most studious of the three crewmen. He realized how fortunate he was to have a second chance, and he was determined to get the most out of the experiments he would operate from the CSM. Like Al Worden, he would be photographing and making measurements over a wide swath of the lunar surface. He spent many hours with Farouk El Baz, Goddard Space Flight Center PIs Isadore ‘‘Izzy’’ Adler and Jack Trombka, and other PIs, learning as much as he could about their experiments and what they hoped to achieve. The results of the Apollo 15 orbital science were now available. Draw­ing on Worden’s experience of operating the experiments for the first time, Mattingly was in a better position to manage them efficiently.

Apollo 16 was launched from KSC just before 1:00 p. m. on April 16, 1972, a more civilized hour for those of us covering the launch. There were no com­plications with the flight until after lunar orbit was achieved, but a major problem surfaced just after the LM and CSM separated. When Mattingly went through his checklist before firing the service module engine to circularize his orbit—his first order of business after separation—one of the gimbal motors that controlled the SM engine nozzle did not respond properly. If he could not get the gimbal motor to work, he would not be permitted to start the engine. Mission rules dictated that the landing would have to be aborted and the LM would rendezvous with the CSM using the LM ascent or descent engine or both. Once joined, LM propulsion would be used to get them out of lunar orbit and on the way back to Earth.

While Mission Control tried to find a solution, the LM and CSM were directed to orbit the Moon near each other but not to join up. The crew, and all of us sitting on the edge of our seats in the SSR, kept hoping the mission would not have to be aborted, but with every orbit that passed without instructions on how to proceed, it was becoming less and less likely that the landing would happen. If the landing was permitted, time lost would have to be deducted from the lunar surface staytime. Eventually, if too much time elapsed, the landing would have to be called off because the sun angle would impair visibility on the surface so Young would be unable to avoid small obstacles at the landing site.

After reviewing the data sent back to Earth from the CSM and consulting with the North American engineers, Mission Control decided it was safe to proceed. Mattingly successfully fired the engine to put him into the desired circular orbit, and Young and Duke completed their landing. The landing delay (approximately 5.75 hours) did reduce the time spent on the lunar surface. To make up for this lost time it was proposed to cancel the third EVA. After much pleading from Muehlberger’s team, emphasizing the importance of the sam­pling sites selected for the third EVA, it proceeded as scheduled but was reduced by about two hours. To keep on the overall flight schedule, the time would be

made up by lifting off from the Moon sooner after the astronauts returned to the LM at the end of the third EVA than originally planned.

With their problems behind them (there were a few others), Young and Duke went about exploring their landing site and deploying all their experi­ments. The ALSEP was set up during the first EVA, and this time the drill worked as designed. But the heat flow experiment met with calamity. After the first hole was drilled and the probe was lowered into the tubular casing, Young, working on another part of the ALSEP deployment, tripped over the cable connecting the probe with the central station and pulled it loose. ALSEP cables consisted of copper wires embedded in a thin plastic covering several inches wide, designed to lie flat on the lunar surface after they were unreeled from their containers. But after being coiled for several weeks in stowage, they tended to develop slight kinks. Whether that caused Young to catch the cable with his boot or whether he just misstepped, the cable came loose. After examining the end of the cable that had been torn off and describing it to Mission Control, it was decided it could not be reconnected.

We never anticipated a failure of this type, so no tools were carried for making a repair. Just to be sure we weren’t overlooking a possible fix, and with a distraught PI begging us to find a solution, we put together another tiger team. But we could not come up with a guaranteed way to reattach the cable. With the cable broken the experiment could not operate, so we canceled the drilling of the second hole. This was a major setback for Langseth; the first deployment of his experiment had placed the probes too shallowly, compromising the read­ings, and this deployment was a complete failure. His experiment, considered one of the most important we would place on the Moon, would have one last chance on Apollo 17. But with the loss of this data point and the compromised data from Apollo 15, even if the Apollo 17 deployment was successful our overall understanding of the heat flow from the Moon’s interior would be open to question, since the measurements at the Apollo 17 site might not be typical for the Moon as a whole.

The deep core was drilled successfully on the first EVA, and eight feet of core were recovered. The active seismic experiment, a duplicate of the one deployed during Apollo 14, functioned much better this time. We had made some minor changes in the thumper firing mechanism, and all twenty-one charges fired. The second part of the experiment, the mortar package, was placed in a better position relative to the ALSEP, and we were able to fire three of the four mortars one month after the astronauts returned home. After firing the third mortar, the pitch-angle sensor showed that the mortar box might have tilted, so we decided to hold off firing the fourth mortar until later.3

The far UV camera-spectrograph, the first telescope to be used on the Moon, was placed in the shadow of the LM and moved several times on succeeding EVAs to keep it in the shadow. It was pointed at different sectors of the sky, three times during the first EVA, four times during the second, and three times during the third, then the film was unloaded and returned to Earth. After setting up the ALSEP and other experiments during the first EVA, the astro­nauts returned to the LM and unloaded the LRV. This left just a short time for the first traverse, and they went to study and sample some craters less than a mile to the west.

On the second EVA the astronauts traveled south about two miles with a major objective of sampling the debris thrown out by a ‘‘recent’’ impact crater (named South Ray) about half a mile in diameter that had scattered ejecta a great distance in all directions. It was expected that this ejecta would provide us with good samples of a geological formation (named Cayley by USGS) that forms extensive highland plains thought by some to be volcanic in origin. If it proved to be volcanic, its composition and age would be important pieces of information for understanding the development of the lunar highlands that make up approximately four-fifths of the Moon’s surface. Equal in importance to resolving the composition of the Cayley formation was obtaining samples of the mountain-making material in the vicinity of the Descartes Mountains. Although the mountains were some distance from the landing site, we believed we stood a good chance of recovering rocks deposited in the plains from impacts that occurred in the highlands.

Tony England, selected in the second scientist-astronaut class, was the mis­sion scientist and CapCom during the EVA periods, and he did a good job of communicating with the crew and relaying questions and suggestions from the SSR. In his role as mission scientist, he had accompanied the crew on many of their training trips and participated in the simulations leading up to the mis­sion, as Joe Allen had for Apollo 15. He was intimately familiar with all the equipment and experiments and was able to quickly give advice when needed.

SSR operations had improved with each flight. Beginning with Apollo 15, we could supply more and more backup information. We kept careful track of the EVAs as they progressed; the planned traverses with each station were identified on a three-dimensional model constructed from the Lunar Orbiter photo­graphs, and we had a list of planned activities for each stop. We also kept a traverse profile showing in graphic form the status of the EVA in terms of time and life-support expendables. Using this profile, we were prepared to suggest modifications to the EVA plans if something unexpected happened or if the astronauts spent more time than planned at a given station. Time was always our enemy, and we knew the crews felt the same. How could they make the most of each minute yet not miss some important discovery? With the pictures coming back from the TV camera carried on the LRV, we were able to keep up with the crews’ efforts and think ahead with them as to what should be the next priority. However, the crews always seemed to make the right decisions without many inputs from the ‘‘back room,’’ a tribute to their training and dedication. There was seldom any second-guessing from those of us privileged to feel so close to the action even though we were 238,000 miles away from actually swinging a hammer or snapping a camera shutter. They were our surrogates in this inhospitable and strange land; we could only sit back and admire the job they were doing. Our job was to try to assimilate the information they were returning so as to arrive at the ‘‘big picture’’ of the Moon that their observations were starting to give us. We were careful not to interrupt the crews or burden them with unneeded questions.

The third EVA that the Field Geology Team had pleaded to retain became a quick trip to sample a northern crater (North Ray), about three miles away. Deposits around this crater were believed by many, but not by all, to be the best opportunity to collect volcanic samples. Time permitted only a few stops, but the traverse was made without incident and harvested many samples. The total weight of samples collected during the three EVAs was 211 pounds, a new record; the total distance traveled, sixteen miles, was slightly less than that recorded for Apollo 15.

Ken Mattingly, after the rocky start caused by the spurious gimbal motor readout, had successfully carried out his part of the science tasks. His sensors and cameras covered much of the Moon’s surface between five degrees north and five degrees south of the lunar equator. Ken also spent as much time as possible making careful visual observations, which were valuable during the crew debriefings and in analyzing the data captured by the sensors.

Several months later, analysis of the Apollo 16 samples showed that they, like most of the samples collected from the previous four missions, were breccias, the products of one or more cataclysmic events that showered the lunar surface with the debris from multiple impacts. None were volcanic in origin, however, and the Lunar Sample Preliminary Examination Team stated that ‘‘no evidence for lava flows or pyroclastic rocks was observed.’’ This was an example of the pitfalls of trying to make photogeologic interpretations with no fieldwork to base them on. But we were learning with each mission how to improve our interpretations, and now we had an additional tool, the CSM data collected from lunar orbit that provided estimates of the composition of the lunar sur­face over wide areas.

With the Apollo 12 ALSEP’s third anniversary of uninterrupted operation approaching, I wrote a memo for Rocco Petrone’s signature that was distributed to NASA management, including the new NASA administrator, James Fletcher, to report how well the ALSEP had performed.4 We reminded those on the distribution that the original design goal was one year and that four experi­ments, the passive seismometer, Suprathermal Ion Detector Experiment, Solar Wind Spectrometer, and dust detector, were still operating normally and re­turning useful data. The Lunar Surface Magnetometer had operated success­fully for two years, and only one experiment, the Cold Cathode Gauge, had failed immediately after activation.

The radioisotope thermoelectric generator was still putting out sixty-nine watts of power, four watts above predicted values for initial power output. The ALSEP central station had responded to over fifteen thousand commands dur­ing the three years and showed no sign of deterioration in spite of having experienced thirty-seven lunations that created temperature swings each time of over 500°F (— 260°F to 270°F; by this time we had a more accurate measure­ment of surface temperatures). The Apollo 12 ALSEP would continue to operate for almost five years longer.

It was 9:00 p. m. on December 6, 1972. Apollo 17, the last of the Apollo lunar flights, was on launch pad 39A at KSC with my good friend Jack Schmitt, Commander Gene Cernan, and CM pilot Ronald Evans strapped on board the command module, named America. As we had done for the last two flights, Rhett Turner and I were prepared to broadcast the launch for the Voice of America Worldwide Service from the press site about three miles west of the pad. To our right, along the raised, curved berm, were the large air-conditioned broadcast booths of CBS, ABC, and NBC, along with booths of other com­panies. If we used our binoculars we could see Walter Cronkite, Jules Bergman, and the other TV commentators looking out their picture windows at the brightly illuminated Saturn Vi Voice of America was a bare-bones operation. We sat in the open swatting mosquitoes, last in line on the berm, on two folding chairs at a card table. Behind us, in what was once a two-wheeled camper trailer, was the engineer with all the electronic equipment. Later, after the lunar landing, along with my regular duties in the science support room at MSC, I was scheduled to make several broadcasts for the Spanish-language VOA using my rusty Colombian Spanish to explain what was happening—my last assign­ment for VOA.

Apollo 17 was scheduled for launch at 9:53 p. m., the first night launch for Apollo. So in addition to our excitement about this last launch and all it meant, we were looking forward to seeing and ‘‘feeling’’ the giant Saturn rocket roar off the pad. We could only guess at the visual impact, the spectacle of the world’s largest firecracker lighting up the night sky. As we sat watching the brilliantly lit launch tower and rocket, the clouds that had partially obscured the sky were slowly dissipating; it had been predicted that if the sky was clear viewers as far as five hundred miles away might see the rocket as it streaked away to the east. Rhett had done his usual impeccable homework for the launch, the best of any reporter I knew, and he carried the audience along as the countdown pro­ceeded, bringing me in as needed to provide some special insight. As we went through our rather informal script, Rhett would signal the engineer to play a previously recorded interview with Jack, Gene, or Ron or some other pertinent clip that would give interesting background about the mission. Illuminated in front of us, between our position and the pad, was the large digital countdown clock counting down the seconds. From time to time Chuck Hollingshead, the voice of Kennedy launch control, would interrupt our coverage with comments broadcast over the public address system.

Everything was proceeding normally until T minus thirty seconds, when without any warning a hold was announced. No immediate reason was given, and we were left, with all the rest of the commentators, to speculate on what was wrong. For the next twenty minutes we tried to make educated guesses, hoping that it was something minor and the countdown would soon resume. Finally Hollingshead came on the PA speaker and explained what caused the hold. The third-stage fuel tanks had not pressurized on schedule, and though a manual pressurization was attempted it was too late in the countdown and the auto­matic sequencer shut down the launch. It didn’t sound serious, but it wasn’t clear when the count would start again; we assumed it might take about an hour before the countdown would resume at T minus twenty-two minutes, the newly announced recycled starting point.

We underestimated. For the next two hours Rhett and I filled the airways with impromptu discussions of Apollo 17 science and whatever other subjects we could think of. From time to time VOA would break away to provide news of the world and return to us, still sitting under the stars waiting for an announce­ment. Listening to those VOA tapes twenty-five years later is a real trip down memory lane. They record the late-breaking news and include the story that former president Harry Truman, age eighty-eight, was in critical condition in a Missouri hospital.

Finally the count resumed at T minus twenty-two minutes; it was held again at a planned point at T minus eight minutes to check that the pressurization trouble had been resolved before the countdown continued. At 12:33 a. m. on December 7, 1972, Apollo 17 was launched; a historic date, one that will always be remembered along with another December 7 thirty-one years earlier. The liftoff was every bit as spectacular as we had hoped, lighting the night sky for miles around and pounding our bodies with the powerful low frequency rever­berations that only a Saturn У launch produced. If you have witnessed a shuttle launch, multiply the effect by two. The crew of Apollo 17 was on its way, and to top it off, we had survived two hours of unscheduled airtime! We packed our notes and left for Houston.

Apollo 17, after its prelaunch difficulties, was the most trouble free of any of the missions. All the pieces were falling neatly into place. We were definitely learning, but now we had no further chance to put this hard-won education to use. The landing site, Taurus-Littrow, almost as far north as the Apollo 15 landing, was on the edge of Mare Serenitatis, to an Earth observer the right edge of the man in the Moon’s right eye. The landing would be the most difficult maneuver of any flight yet, requiring Cernan to come in over the Taurus Moun­tains, 6,500 feet high, descend steeply into a narrow valley, and land between the bases of two mountains.

Taurus-Littrow was selected for the final Apollo landing for several reasons. During the Apollo 15 mission Al Worden had observed that this area was covered by a mantle that looked darker than other parts of the lunar surface. His observations seemed to be confirmed by Lunar Orbiter photography and photographs taken from orbit during other Apollo missions. The promise of finding “recent” volcanism raised its head again. Would this site provide sam­ples that would confirm an epoch of late lunar volcanic activity? Samples from the Taurus Mountains were also of great interest. Would they be similar to or different from the highlands samples collected on Apollo 16?

As a far northern and eastern landing site, it had value for several of the ALSEP experiments, in particular for the Lunar Atmospheric Composition (LACE) experiment (the passive seismometer was not carried on this mission), which would provide better data if separated in distance from the other ALSEPs, which were still sending measurements. LACE, a miniature mass spectrometer, was a more sophisticated version of the Cold Cathode Gauge experiment de­ployed on missions 12, 14, and 15 to detect the tenuous lunar atmosphere. We also had two new surface experiments on the mission, Surface Electrical Proper­ties (SEP) and the traverse gravimeter, which, along with the portable magne­tometer that flew on Apollo 16, would be operated by the astronauts during the LRV traverses. These last three experiments were expected to provide important information on the subsurface structure of the valley at the base of the Taurus Mountains.

While missions were under way, my job at MSC included manning a console in the Science Support Room and taking part in the discussions that would fill the exciting hours while the astronauts were on the lunar surface. Occasionally I would spell the headquarters duty officer in the Mission Operations Control Room, the latter largely a ceremonial duty if the mission was proceeding ac­cording to plan. In addition, I would participate in briefings with VOA and other news organizations. Although the word ‘‘spin’’ had yet to be applied to government briefings, that was part of our approach. If something in the mission timeline didn’t go according to the material passed out to the media before the mission, we would be interrogated at each of the daily updates held in the MSC auditorium. No question was off limits, and some would be off the wall, reflecting the media’s understanding, or misunderstanding, of what was going on.

We all had our preferred media person to talk to off line, someone we knew from experience would tell the story reasonably straight and get the facts right. My favorite was Donald Kirkman of Scripps-Howard. I tried to avoid Thomas O’Toole of the Washington Post, with some success, for he seemed to be always looking for the negative side of events and could usually be counted on, at some point during a mission, to misinterpret an important story. We would feed trusted reporters tidbits of insider information so that their stories would be more informative or have a little more punch than their competitors’.

Just before the crew achieved lunar orbit, their discarded SIVB stage hit the Moon about 525 miles west of the Apollo 16 ALSEP, and the impact was re­corded by all four passive seismometers that were still operating from the earlier missions. This time there were no problems in lunar orbit after separa­tion of the LM and CSM, and Cernan accomplished the landing after taking over control from the autopilot and set the lunar module, Challenger, down in the rock-strewn valley between the North and South Massifs.

Once on the surface, Cernan and Schmitt, the last men to set foot on the Moon for what has turned out to be three decades and counting, went energet­ically about their business. The crew had trained hard, and we could sense from Cernan’s descriptions that he was not about to be outshone by his geologist teammate when it came to conducting their surface studies. They described the sight that confronted them as “spectacular.” I have probably overused that word in this story as much as the astronauts did during the missions, but it is the best adjective I know to describe the views we could see from the TV images and later from the many excellent photos they returned. To the north, less than seven miles away, the mountains rose almost perpendicular from their base toward the black sky. To the southwest, again less than seven miles away, lay the South Massif, equally imposing if not quite as steep as the mountains to the north. TV pictures captured the landscape clearly, and in the SSR we could only wonder at our audacity in asking the crew to land in such close quarters.

Apollo 17 reconfirmed the targeting ability of the MSC engineers. They brought Cernan and Schmitt to the precise point where Cernan was scheduled to take control, and he then successfully demonstrated his landing skills. With this experience it seems certain that if missions had been scheduled after Apollo 17 we could have persuaded management to agree to landings at important sites such as the central peaks or rims of Copernicus and Tycho. Future lunar explorers, undoubtedly piloting spacecraft with greater capabilities, will find safe landing sites almost anywhere on the Moon, including the farside!

Apollo 17’s first EVA began with the removal of the LRV from its stowage bay on the descent stage and the erection of the TV high gain antenna. Thereafter we had good TV coverage of the landing site and the astronauts deploying the ALSEP. They drilled three holes, two for the heat flow experiment and the third to recover a ten-foot core. ALSEP began transmitting data as soon as it was activated, and the star of the experiments, the Surface Gravimeter, provided strong signals. Little did we know at this point that there was a major problem, as described in chapter 7. After finishing these tasks, the astronauts still had time to take a short (about half a mile) ride to the south to collect samples near a small crater named Steno. During this traverse they also took Traverse Grav­imeter and SEP readings and left two explosive charges to be detonated later to provide signals for the Seismic Profiling experiment.

A few words about the explosive packages that were an integral part of the Seismic Profiling experiment. Commencing with the design of the active seis­mic experiment, carried first on Apollo 14, we went through an extensive review and certification of the explosives used with that experiment and the Seismic Profiling experiment. Some at NASA were not happy about carrying live explo­sives on the LM, so our test procedures were carefully monitored. We had to prove beyond any doubt that there could be no accidental firing of the charges. Petrone, especially, followed the certification process from beginning to end and witnessed some of the field tests.

Fortunately this experiment was not the only place explosives were used during the mission, starting with the separation of the launch escape tower from the CSM and progressing through the individual rocket stages, where explosive squibs were used to separate some of the stages during flight. We benefited from all the work that went into qualifying these explosives and designed our charges using aspects of these proven designs. The biggest fear, of course, was that an inadvertent firing command, short circuit, or other acci­dent might trigger the explosives, either while they were stowed on the LM or while the astronauts were setting up the experiments.

Because the astronauts would hand carry the explosives on the lunar surface, every firing circuit had either double or triple safety redundancy before the firing commands could activate the charges. For the Seismic Profiling experi­ment, the arming sequence was as follows: Each explosive package had three pull rings on top. Pulling ring one started the safe/arm timer. Pulling ring two, and rotating it ninety degrees, released the safe/arm slide to start the mechan­ical timer. Pulling ring three cleared the firing pin and placed a thermal battery timer on standby until a coded signal was received from the ALSEP central station, the preferred way to set off the charges. In case ALSEP commands weren’t received, the mechanical timers were preset for periods from 89.75 to 92.75 hours after activation, well after the astronauts left the lunar surface. Each charge package had an antenna that would receive the initiation signal from the central station to start the firing sequence, at which point there was a two – minute window in which to receive the coded firing signal.

Sound complicated? It was. There was a running joke that with all the safety features we would be lucky to get even one to fire. But all eight charges were fired successfully after the astronauts departed, and the experiment’s four geophones recorded the explosions, providing information about the upper mile and a half of the Moon’s subsurface. The LM ascent stage impact, about seven miles southwest of the landing site, with its higher energy input, allowed Bob Kovach, the PI, to improve his measurements and estimates of seismic velocities to a depth of three miles below the surface.

The second EVA, the longest of the three, went almost due west and then swung southwest to study and sample the South Massif. Cernan and Schmitt made several stops along the way, including sampling the rim of a fairly large crater called Camelot. On finishing their work at the crater, they drove up the low scarp that separated the valley from the massif and sampled the boulders at the base of the massif. They then returned to the LM by a different route that took them farther north on the valley floor and included seven more sampling stops. One of these stops, at Shorty Crater, was to sample the dark halo material surrounding the crater that could be seen on Lunar Orbiter photographs. Worden had reported that he could see a color difference from orbit. Once again we hoped the dark material would be ‘‘recent’’ volcanic deposits as was predicted (incorrectly) at the Apollo 16 landing site. What they found, to the great excitement of both astronauts, was an orange and red soil interspersed with darker and lighter soils. Schmitt thought they had found the elusive recent volcanic vent. Returning to the LM at the end of the EVA with their find, they placed three Seismic Profiling explosive packages at varying distances from the ALSEP central station and took a series of Traverse Gravimeter and SEP readings.

The third EVA traverse was made toward the east and then turned north to sample the North Massif and the intervening darker plains material, also inter­preted as volcanic mantling material. Traverse Gravimeter readings were made on this EVA, and the final three Seismic Profiling explosive charges were placed on the surface at intervals along the route. Because the SEP receiver overheated, no data were collected by this experiment during the third EVA.

By now, after analyzing the signals coming back for almost two days, we realized that the Surface Gravimeter was not responding correctly. We cut the third EVA short to allow Schmitt to go back and rebalance the gravimeter’s movable beam. This last attempt to improve the experiment’s response while the astronauts were still on the Moon also failed. Something was wrong, but we weren’t sure what it was and couldn’t find a solution. Closing out the final Apollo lunar surface EVA on a somewhat dismal note because of the gravimeter problem, Cernan removed the neutron probe from the core hole for analysis back on Earth and climbed back into the LM.

With the crew of Apollo 17 safely on the Moon, Dale Myers and Rocco Petrone released the final ‘‘Apollo Program Plan.’’5 It covered all the remaining Apollo 17 activities and those actions necessary to close out the Apollo program. Among other notes of finality, it stated: ‘‘All basic hardware procurement for the Apollo Program has been accomplished.’’ The nation would not purchase any more awe-inspiring Saturn Vs or superbly engineered CSMs and LMs. The schedule in the plan showed a transfer of responsibility for common Apollo – Skylab activities to the Skylab Program Office, the next approved program, by the middle of FY 1973. Beginning on the same date, the remaining Apollo lunar science activities, mostly monitoring the ALSEPs and publishing results, would be undertaken by the Office of Space Science and Applications. Although Apollo 17, and for that matter the entire Apollo program, had achieved all its objectives and more, this final Apollo plan ended an era with the sour taste of a great opportunity lost through lack of national leadership. It was an era that had begun with great expectations of conquering new worlds.

After almost seventy-five hours on the lunar surface, twenty-three of them spent outside the LM on the three EVAs (another new record), Cernan and Schmitt lifted off to rendezvous with Ron Evans, carrying with them the 243 pounds of samples they had collected at Taurus-Littrow during traverses that covered more than twenty-one miles. The LM was jettisoned, and this time the impact occurred just west of the landing site, some 475 miles east of the Apollo 15 ALSEP. Its impact was recorded by all four of the previously deployed passive seismometers and the Seismic Profiling experiment deployed on this mission.

While Cernan and Schmitt were on the surface, Ron Evans had been con­ducting the experiments assigned to him. He had also undertaken some experi­ments beginning with the translunar coast phase and would continue making measurements during the return home, almost until reentry. The high activity period, however, was while he was in lunar orbit, where he conducted a suite of experiments similar to those of the previous two missions. Once Cernan and Schmitt were on board and they were on their way back to Earth, Evans carried out an in-flight EVA when he retrieved the film canisters and other data from the SIM bay. Splashdown and recovery were uneventful.

Project Apollo was now a new chapter in the history books. Even with a few glitches, the flight of Apollo 17 had been the most successful mission from a scientific viewpoint. An enormous treasure trove of lunar samples was in the vaults at the Lunar Receiving Laboratory awaiting study. The seismic and laser corner reflector networks were already returning exciting information, as were many other ALSEP experiments. ALSEP central stations were performing up to or beyond their original design goals. But political and public apathy had set in long before the launch of Apollo 17, and the scientific results alone couldn’t convince the decision makers to add more missions. Those of us still working in the Apollo Program Office faced the dismal task of mopping up and closing down an unequaled undertaking. Many of my coworkers had already begun to drift away to other NASA offices or to new work in or outside the government.

Our dreams for lunar exploration never went away; we always hoped that Congress and the Nixon administration would see the error of their ways and provide the funding to reinstate our post-Apollo plans. But in spite of the surprising discoveries made by Apollo 11 and the missions that followed, no national commitment was forthcoming, and the Apollo hardware remaining after Apollo 17 was never used for its original purpose. Some was used for Skylab, some for Apollo-Soyuz, and the other items are lying ignominiously on the ground at museums, like tethered Gullivers, as reminders to millions of visitors each year of Apollo’s magnitude—and perhaps, to some, of oppor­tunities lost. When will man again set foot on the Moon? Or will we bypass the Moon and go directly to Mars? Or will we stay earthbound or in near-Earth orbit for generations?

One afternoon, walking between my office at L’Enfant Plaza and the NASA of­fices at 600 Independence Avenue, I met some of my former Advanced Manned Missions and Apollo colleagues who had recently been assigned to potential future manned space flight programs. We talked briefly about the uncertainties surrounding these programs (none had been officially blessed as the successor to Apollo except for the short-term Skylab and Apollo-Soyuz programs) and discussed where I might find a new job. I was struck by their lack of enthusiasm and their pessimism about their new work as we discussed NASA’s future. It seemed as if almost overnight this marvelous can-do agency had grown old and lost its way. Gray heads were beginning to predominate in all the offices. Instead of looking toward an exciting future, everyone seemed to be scrambling to hang on and find a place to roost. I decided at that moment that it was time to leave NASA and find a new program I could devote my energies to.

Following the lead of Ed Davin and Dick Green, I sent an application to the National Science Foundation and was hired immediately by a new organization called Research Applied to National Needs, which was undertaking research on a wide spectrum of new technologies. Thus began a new career, but never again would I experience the excitement and the sense of achievement that came with being a small part of Project Apollo.

The Legacy of Apollo

One of the questions most frequently asked at the end of Project Apollo, and even today, is What did we learn? It’s a good question. It would often be followed by other, unanswerable questions. Was the project worth the cost? Wouldn’t we have been better off spending the billions of dollars on X, Y, or Z?

Addressing the unanswerable questions first from the perspective of science, it is difficult to calculate the part of Apollo’s cost that funded scientific experi­ments for the Apollo flights, because the many components that made up Apollo science were carried in different parts of the NASA budget. Should it include the salaries and overhead for all the civil servants involved? Should it include the support contractors’ costs for those who worked at NASA centers and headquarters and were involved in the planning and development of the science? How about facilities such as the Lunar Receiving Laboratory? The LRL cost over $16 million to build and equip, plus additional operating expenses during the missions. Add to that sum the $19 million given to the sample analysis principal investigators, and the expenses for quarantine and sample analysis alone total over $35 million.

Considering how the post-Apollo studies contributed to Apollo science, should any of those costs be added to the total? Should all the advisory commit­tees and summer conferences that were funded by NASA? They contributed important advice and helped us select the experiments included on the mis­sions. And of course there were the costs associated with integrating experi­ments on the lunar module and the command and service module and training the astronauts in their use and deployment. Finally, there is the cost associated with experiment data reduction. Calculating an accurate sum for all these activities is probably impossible at this late date, and the items mentioned probably overlook other costs that would contribute to a grand total.

NASA bookkeeping, like that of many government agencies and cabinet departments, used a document called a program operating plan. The POPs categorized expenditures by program, and within each program the expendi­tures were further delineated by a work breakdown structure or, in simpler language, an item-by-item accounting. These terms are important only to show that there was great rigor in keeping track of taxpayers’ dollars. Each office and center within NASA kept these records, and they were compiled and reviewed by the NASA headquarters Office of Programing. This office not only kept track of expenditures but was also the focal point for preparing each year’s budget requests to the Bureau of the Budget and its successor, the Office of Manage­ment and Budget, and then with other senior management presenting and defending the budget before Congress. The name of this office changed through time, but the men who ran it, such as DeMarquis ‘‘Dee’’ Wyatt and William Lilly, were both feared and admired because of their power to approve or disapprove program requests.

In the Apollo era, budget control was a hallmark of NASA, and discussion of the flow of funds for all programs took up a major part of Jim Webb’s and Bob Seamans’s monthly program reviews with the associate administrators and lesser managers. Program managers were expected to keep their books up to date and in good order. Any deviations from approved schedules and budgets during an individual program’s lifetime had to be fully explained and justified, on pain of strong reprimand or even demotion or removal. Considering the uniqueness of this new frontier and the challenges it represented, only a few large overruns occurred. The Surveyor program was an example. The problems that have plagued the International Space Station in recent years, including schedule delays and large cost overruns, would not have been tolerated in the early days of NASA by either NASA management or Congress. But that is a story for another day.

My ‘‘hard’’ number for Apollo science includes estimates of the components listed above and is based on reviews of microfiche records, internal memos, and POPs, of which the last one I had access to was POP 72-1C.1 There may have been later POPs covering Apollo science, but this one showed closeout costs for the last four years of the surface and orbital science programs and contractor manpower ramping down toward zero. The total reported in POP 72-1C was $150,000,000—a nice round number, but I believe it was understated. In a memo to NASA Public Affairs, responding to a request for the costs of the Apollo 15 experiments, we made an estimate of $36 million.2 In another memo, this time to the NASA budget office in March 1972, we estimated a total expenditure of $85 million for the Apollo 16 and Apollo 17 missions, a total of $121 million for just the J missions.3 This indicates to me that the 1972 POP did not include important pieces; however, you can’t tell the basis for the number— what was included or what may have been left out. In William David Comp­ton’s history of Apollo, Where No Man Has Gone Before, he indicates that slightly more than $218 million was spent on science payloads.4 But once again it is not completely clear what this number represents.

An estimate I made at the end of the program was $265 million, based on calculating the cost of each experiment and adding other related contractor costs available at the time. But that estimate did not include some of the items described earlier. I now believe the total would come close to $350 million in 1972 dollars, not including civil service salaries and benefits. If you accept this number, the science piece of Apollo was about 1.5 percent of the total $25 billion spent.

If we factor into the $25 billion the national prestige value of being the premier spacefaring nation, the excitement of visiting a new world, the knowl­edge gained about the Moon and Earth, and the advanced technologies that resulted from Apollo (some call it spin-off), we can try to answer the unanswer­able. Was it a bargain, money well spent, or money wasted? My judgment: unequivocally a bargain!

What did we learn? remains an important question because as students of the Moon continue to examine material brought back during the Apollo mis­sions, fresh results are still coming in. New information from the recently completed Clementine and Lunar Prospector missions adds to our knowledge and clarifies or extends the Apollo results. More than 1,100 abstracts were received for the Thirtieth Lunar and Planetary Science Conference held in Houston in 1999, approximately one-quarter dealing with lunar subjects, at­testing to the continued interest in Moon-related studies. Programs to return to the Moon, based on a desire to learn more about our nearest celestial neighbor and perhaps begin to exploit its resources, are constantly proposed.

If, as many of us who worked on Apollo fervently hope, the United States (perhaps in concert with other nations) mounts another Apollo-type project to send astronauts back to the Moon or on to Mars, then we must be prepared to justify and explain to the American public the benefits of spending a nontrivial amount of the national budget on such undertakings. At the moment NASA management does not support going back to the Moon, on manned or un­manned missions. In spite of the interest in recent Mars missions, sustaining public support for extended Mars exploration will be difficult. To the casual observer, or the average taxpayer, one picture of a Mars landscape will look much like the last one, even if it includes an astronaut holding a rock, pointing at a mountain, or riding around on some strange-looking vehicle.

If a political objective is not the driving force at the time the debate begins, as it was at the start of the Project Apollo, then we must be able to predict scientific and economic benefits of value to those on Earth. Such predictions will be difficult to make unless we can provide a connection to what we gained from the Apollo Moon landings and extrapolate this knowledge in a rational way to these new projects. Intellectual adventures will not suffice, even if one believes that the thrill of exploring new lands still survives in our species.

Briefly, here is a summary of the findings and lessons learned from Apollo and the Lunar Orbiter missions that immediately preceded Apollo. A few of Lunar Orbiter’s contributions are discussed briefly, and the notes list references that provide details on this program. From all the missions and other pro­grams, such as the radar studies conducted from Arecibo in Puerto Rico, by the end of Apollo we came to a new understanding of the Moon.

Lunar Orbiter’s comprehensive, high resolution coverage of the Moon’s sur­face allowed lunar students to expand their understanding of the Moon in significant ways.5 For example, the higher resolution pictures permitted the United States Geological Survey lunar mappers to refine the geological studies they had been making for the previous four or five years based on telescopic observations. Before Lunar Orbiter returned its magnificent photographs, geological formations mapped by USGS workers were distinguished by such characteristics as subtle differences in albedo (reflective power), surface rough­ness, and crater counts. With higher resolution Lunar Orbiter photographs in hand, the quality and speed of their work increased. The validity of their interpretations would have to await the additional information to be returned by Apollo missions.

In retrospect, one would have to give USGS a good grade (perhaps a B+) for its early efforts. Physical differences were no doubt present; what they repre­sented was difficult to predict. Forced to make interpretations based on these subtle distinctions, and working under the great disadvantage of not having material in hand that represented their mapped formations, some overesti­mated the complexity of the Moon’s surface. Perhaps the best illustration was the view that many places on the Moon exhibited volcanic features such as cinder cones. The Apollo 16 landing site, selected in part to permit the astro­nauts to sample this type of feature, returned mostly breccias and no volcanic ejecta. However, the famous ‘‘orange soil’’ found at the Apollo 17 site is inter­preted to mean that it was formed during lava fountaining from a volcanic vent, but almost 4 billion years ago. No traces of ‘‘recent’’ volcanism were found. Nothing significant seems to have occurred on the Moon for at least the past 50 to 100 million years except for random impacts.

Lunar Orbiter’s farside coverage allowed the USGS mappers to extrapolate their extensive nearside studies to this perpetually hidden face. Its appearance, highly cratered and without the vast, smooth maria common on the nearside, differed from the face of the Moon that had been studied for centuries. It looked much more like the nearside lunar highlands. With a few exceptions, such as the large crater named Tsiolkovsky and the Mare Moscoviense basin, the large farside impacts had not filled with marelike flows as had many of the large nearside impacts. This difference was attributed to the pull of Earth’s gravity, with the nearside being much more strongly influenced during the early history of the Moon than the farside, thus allowing lunar basalts to fill the low – lying nearside basins. This conclusion supported the belief that early in its formation the Moon had become locked into its present orbit, with its orbital rotation around its axis of twenty-eight days equaling its orbital period around the Earth.

Lunar Orbiter also permitted a more detailed analysis of the Moon’s gravita­tional field and the irregularities in the field. Its ability to provide this informa­tion had been proposed by Gordon McDonald. By closely tracking each space­craft’s orbital path and calculating how it differed from the path that would be expected if the Moon’s gravity field were uniform, lunar geodesists were able to accurately plot, for the first time, the figure of the Moon. This close tracking led to the discovery of the ‘‘mascons’’ (mass concentrations) mentioned earlier. Deviations from Lunar Orbiter’s calculated flight path suggested that material denser than the surrounding terrain formed the widely scattered mascons. These data were upgraded by tracking the CSM on each Apollo mission, and more recently they have been refined by tracking the orbits of Clementine and Lunar Prospector.

Finding the mascons has important geological and geophysical implications that should contribute to deciphering the Moon’s early history. In addition, knowing the mascons’ positions will be especially useful when we return to the Moon with either manned or unmanned missions, because it will allow us to program the landers to arrive precisely at their designated landing sites. But Lunar Orbiter, Apollo, and recent missions could not tell us what the mascons are or how they were formed. Resolving these questions will have to await additional measurements made by the next generation of spacecraft.

Ranger, Surveyor, Lunar Orbiter, and Apollo put to rest for most lunar students the question of the origin of almost all lunar craters: they were formed by impacts. This knowledge has led us to look at the Earth’s history in a new light.6 Before the Apollo landings, most Earth scientists believed that in its earliest history the Earth had witnessed a period of intense infall of large and small planetesimals, meteorites, and other debris from a newly forming solar system. Little direct evidence of this epoch could be found in the geological record, and until recently only a few impacts had been positively identified and studied. The rate at which these large and small impacts bombarded the Earth was pure speculation, but those who studied these features believed large im­pacts were probably common.

Today some terrestrial features and events that had previously been difficult to explain are being attributed to large impacts. The most fully reported event of this type provides an explanation for the disappearance of the dinosaurs and many other species of animals and plants at the end of the Cretaceous period, some 65 million years ago. Proposed in 1980 by Luis Alvarez, his son, and several other researchers,7 the theory was based on the discovery in Italy of a thin rock formation, enriched with the element iridium, at the Cretaceous/ Tertiary geological boundary.

They concluded that the best explanation for this anomaly was that a large object composed of material containing a high percentage of iridium, a com­mon constituent of certain types of meteorites but not common in Earth rocks, had struck the Earth at this precise time. Debris from this impact spread over a large portion of the Earth’s surface and was deposited as a thin layer that included the formation discovered in Italy. This proposal was met with great skepticism by many in the scientific community, but some, including Gene

Shoemaker and others who had been involved in Apollo science, supported the idea, knowing that large impacts had affected the Moon’s history. This Creta­ceous impact has been confirmed, and through the work of many scientists, the probable impact site has now been located on the edge of the Yucatan peninsula in the Gulf of Mexico. Whether it led to the species extinctions observed at the end of the Cretaceous period is still being debated.

The Apollo program’s emphasis on understanding impact craters spurred the search for and discovery of other large Earth impacts. For example, an ancient impact crater has been found in Texas, at Sierra Madera; another underlies Chesapeake Bay; and a buried crater in southeastern Nevada is be­lieved to have created the Alamo breccias. The identification of impact events in the geologic past has accelerated as our diagnostic techniques have improved. Australia has been especially productive for the study of impact craters be­cause much of its surface has remained relatively undisturbed for millions of years. It was while undertaking such a study that Gene Shoemaker met his untimely death.

These discoveries have led to a related field of study, tracking objects orbiting near the Earth and crossing the Earth’s orbit (hundreds are now known) that might strike the Earth in the future. Today, if such an object took aim at Earth, it could not be avoided. If another object the size of the one that hit the Earth at the end of the Cretaceous period were to strike the planet, it would trigger a series of events with unimaginable consequences. But not much is being done to prepare for such an admittedly low-probability event. Some believe we could avoid such an impact, if it was predicted, by developing an early warning system that would track large meteors or asteroids and then deflect them with missiles. The study of impacts on the Earth and Moon has resulted in a model that predicts the frequency of impacts on the Earth. This model suggests that a large impact occurs approximately once every 50,000 to 100,000 years. Perhaps this knowledge will motivate world governments to work together for a solution that will prevent such a catastrophe.

Although a relatively small event when it occurred, the Meteor Crater im­pact has been dated at approximately 50,000 years ago. It undoubtedly was a devastating blow for a large region surrounding the impact point, creating ground tremors and clouds of dust and debris that would have extended over hundreds of square miles. At that time the only casualties may have been a few mastodons and other wildlife. If such an event occurred today, Flagstaff and other nearby towns would probably be destroyed, and cities as far away as Tucson and Phoenix would feel its effects. Are we due for another big impact- soon? The model suggests we may be.

By the end of the Apollo missions, the six successful landings and their predecessors had returned a wealth of new information about the Moon. Before the landing missions, Apollo 8 and Apollo 10 traveled to the Moon but did not land. Apollo 8 was the historic mission that orbited the Moon at Christmas 1968, with men being captured for the first time by the gravity field of a ‘‘planet’’ other than the Earth. Although we had in hand excellent close-up photographs from Lunar Orbiter, this was the first time men were able to view the Moon at close range.

The lunar farside especially impressed the crew of Apollo 8; Frank Borman, Jim Lovell, and William Anders reported seeing a jumble of craters on top of craters. Orbiting sixty-nine miles above the surface, they described the Moon during their Christmas Eve greeting to those back on Earth as ‘‘a vast, lonely, forbidding. . . expanse, . . . it certainly would not appear to be a very inviting place to live or work.’’8 Fortunately we were not going to try to make an Apollo landing on that ‘‘forbidding’’ farside terrain. Apollo 8 also gave us the first views of our home planet from a great distance away; the Earth was described as an oasis, isolated in the emptiness of space. Some have credited this dramatic view of Earth with imparting a new awareness of how unique our planet is and how important it is to protect its fragile environment-an unexpected bonus from the Apollo program.

Apollo 9, launched in March 1969, was the first test of all the Apollo hardware working together as it would for a Moon mission except that the crew and the equipment never left Earth orbit. It was followed two months later by Apollo 10, a dress rehearsal for the first attempt to land on the Moon. The crew of Tom Stafford, Gene Cernan, and John Young would perform all the complicated maneuvers required of a landing mission except for the most crucial-the actual landing. Stafford and Cernan would separate from the CSM in lunar orbit, descend to less than ten miles above the lunar surface, jettison the landing stage, activate the LM ascent engine, and rendezvous with the CSM. Close, but oh so far from making history. In addition to testing all the elements leading up to a landing, they proved the accuracy of Apollo targeting and the astronauts’ ability to see their landing point and observe potential hazards at a site similar to that expected for the first landing in the Sea of Tranquility. Apollo 8 and

Apollo 10’s reconnaissance also confirmed what had been seen in the Lunar Orbiter photographs: smooth landing areas were available in the ‘‘Apollo zone.’’

Cernan and Young would get another chance to perform a Moon landing; Stafford is the only man to get within ten miles of the Moon and never land. I’m sure he would gladly forgo that honor for the thrill of having kicked a little Moon dust. Two months later the first landing would take place. Many other Apollo prime and backup crew members—Walter Schirra, Donn Eisele, Walt Cunningham, Jim McDivitt, Gordon Cooper, Joe Engel, and Rusty Schweick- art—would suffer the disappointment of being selected to test the Apollo hard­ware but never getting to the Moon. But without their key roles and dedication the Moon landings could not have been undertaken.

By the time Apollo 11 splashed down we had developed the routine by which the science results would be processed and disseminated. The astronauts would be picked up by a navy helicopter operating from an aircraft carrier, transferred to a specially designed trailer on the carrier, and flown back to Houston to be placed in quarantine in the LRL. The samples, film, and other data would be removed from the command module and flown to the LRL in their own air­craft. Once in the LRL, the astronauts would be debriefed by a team of scientists and engineers while the samples were unpacked, examined, and cataloged and the photographs were developed. In the meantime, we would be receiving data from the instruments left on the Moon.

This routine was followed for all the missions, with the major difference that after Apollo 14 the astronauts and the samples no longer had to spend time in quarantine and the debriefings became much more relaxed and easier to carry out. Without an intervening barrier, we could question the crews much more directly as we tried to piece together all they had done. This was an important change, because the last three missions were more complex and the astronauts’ recollections more valuable in reconstructing their long traverses on the lunar roving vehicles. I will describe only how we debriefed the Apollo 11 crew and studied the first samples, but I will include results from all the missions to explain what Apollo taught us.

The Apollo 11 astronauts had their first science debriefing on August 6, 1969. Before that debriefing the Manned Spacecraft Center engineers had reviewed the nuts and bolts of the mission—the ‘‘technical debriefings’’—going over those aspects of their flight that might affect the success of the next mission. Although all spacecraft systems were carefully monitored by telemetry, with records kept of all discrepancies, the astronauts’ answers to questions would often clear up troubling inconsistencies or uncertainties in the records. Neil Armstrong’s and Buzz Aldrin’s descriptions of their landing maneuvers and their difficulties in finding a good landing site were examples of how their experience contributed to improving the landing sequence for Apollo 12 and later missions. All members of the crew of Apollo 12—Pete Conrad, Dick Gor­don, and Al Bean—as well as the backup crew led by Dave Scott, were the most interested participants in these debriefings. Pete must have gained valuable knowledge, because he landed right on the money, within easy walking distance of his target, Surveyor 3.

Although I was invited to the science debriefing, I sat in one of the back rows while a few designated individuals, including Don Wise and O. B. O’Bryant from our headquarters office, were allowed to ask questions. It was a strange scene for such a momentous occasion, with the questioners and hangers-on peering at the three astronauts, who sat behind a brightly lit picture window like animals in a zoo. Unfortunately the transcript of the debriefing does not always identify the questioner, but Gordon Swann and Henry Holt of USGS and MSC’s Gene Simmons, among others, covered all the important questions relating to the astronauts’ surface observations, especially those that might affect what was planned for the next mission.9

Everyone involved in this debriefing, and in debriefings for later missions, came away with a great admiration for the astronauts’ powers of observation and recall. When these traits were added to their innate resourcefulness and doggedness in following and going beyond their ambitious timelines, every possible ounce of science was gleaned from the missions in spite of the con­straints they were working under. Some might take issue with that statement, but I believe it is true; the training and simulation had paid off beyond our expectations. Explorers of all generations have been eulogized for daring to take chances beyond the imagination of the ordinary person—for the astronauts it was called ‘‘the right stuff.’’ However you wish to identify this urge to explore, it was undeniably present in these first voyagers beyond the friendly Earth who risked never returning to their home planet, a danger never before faced by explorers.

The Apollo 11 science debriefing was our first chance to talk directly to the crew after their return from the Moon. After two weeks of isolation, interro­gated every day by engineers and technicians, the astronauts were in a surpris­ingly good mood. From time to time one could sense a little irritation at questions that were repetitive or trivial, but all in all there was great coopera­tion, and we gained much from listening to their firsthand observations while they were still fresh in their minds. Their photographs had been developed and were available to supplement the discussion, as was the preliminary traverse map of the landing site. When necessary, the astronauts used large pads and marker pens to illustrate their answers, and Armstrong, especially, took advan­tage of these aids. (Were these unique drawings preserved for future genera­tions of historians?)

One of many exchanges was particularly interesting. While in lunar orbit, before beginning their return to Earth, the astronauts were asked to look to­ward the crater Aristarchus and describe it. Although Aristarchus was just on the horizon and at the limit of their view, Armstrong reported that he thought he saw fluorescence in that region. This announcement caused some stir: Was he observing some lunar transient phenomenon like that described in chapter 2? Now, during the debriefing, he went into more detail and modified his ob­servation. Although he described the general area as the brightest spot he could see, he could not confirm that it was Aristarchus itself that was causing the bright reflection, and he did not ‘‘mean to imply that it was self-illuminated.’’ The unusually bright appearance of the Aristarchus region to the crew in orbit reinforced the belief that it might be the site of recent activity on the lunar surface. With their many other observations and much hard work, the crew of Apollo 11 had opened the door to a new era in planetary science.

The possibility of bringing some deadly unknown disease to Earth in the samples or by an infected crewman led to building the Lunar Receiving Labora­tory. Strict protocols had been developed to guard against these risks. Some in the media latched on to this potential hazard, attempting to fan the public’s fears of some catastrophic invasion. One month before the Apollo 11 liftoff, a media briefing was held in Washington to describe the details of the mission and, we hoped, allay any fears that the first landing and return from the Moon’s surface posed any danger to life on Earth.10 The final portion of the briefing was conducted by Air Force colonel John Pickering, who had served on the Inter­agency Committee on Back Contamination and now held the title director of lunar receiving operations at the Office of Manned Space Flight. He went to great lengths to describe the procedures that would be followed, from collecting and packaging the samples on the Moon through recovery and transport of the

samples and astronauts to the LRL and eventual release of both at the end of the quarantine period. He even went so far as to include in the press handout a copy of the LRL biological certification signed by Dr. David J. Sencer of the United States Public Health Service, chairman of the Interagency Committee, to prove that all precautions had been taken. This openness and attention to detail defused this issue for most of the media, and it never surfaced again as a major public concern. However, managers of Mars missions that will return samples to Earth should head off the potential negative exploitation of this issue by being open and detailing the steps that will be taken to guard against alien organisms.

Let me illustrate how seriously the quarantine protocols were followed. During the preliminary study of the Apollo 11 samples, a technician was cata­loging a sample in an isolation chamber glove box that operated under negative atmospheric pressure to avoid any leakage into the LRL when one of his gloves ruptured, exposing him to the sample. This man and another working near him were immediately placed in quarantine in the LRL with the three astro­nauts. One lemon-sized rock was carefully sterilized and taken out of the isola­tion chamber and given to the Lunar Sample Preliminary Examination Team (LSPET) so they could hold it. Cliff Frondel, a member of the team, was quoted as saying, ‘‘It was a great thing to look at this stuff that people had been

speculating about for millennia, and here it was in our hands______ It was a hell of

a thrill.’’11

To determine if there were possible life-threatening forms (“replicating spe­cies’’) in the samples, ten species of animals were exposed to lunar material for twenty-eight days, either through inoculation or in their food. Four control groups were exposed in a similar fashion to nonlunar material for the same period. These animals included paramecia, planarians, shrimps, oysters, cock­roaches, and houseflies. One might wonder if the testers, fourteen scientists called the Lower Animal Test Team, had any second thoughts about including cockroaches, insects that seem to be indestructible and have survived 200 mil­lion years of evolution essentially unchanged. Why would a little Moon dust hurt them, regardless of what it contained? The cockroaches and the astronauts cooped up together in the LRL became the basis of many jokes.

During the quarantine period, these ten species, living in small aquariums or jars and bowls inside the LRL, were carefully monitored for any suspicious behavior or a sudden desire to go to the special heaven reserved for them.

Nothing much unusual was observed; only the oysters, both those exposed to lunar material and the control groups, seemed to have a higher than expected death rate. This was attributed to conducting the tests during their normal spawning season, which apparently is stressful to romantic oysters. But as in all true scientific inquiry, one strange behavior was noted: planarians exposed to heat-sterilized lunar material swam at the surface of their bowl more frequently than the control groups.12 The reason was unknown.

In addition to the study of ‘‘lower animals,’’ similar tests were conducted on mice and quail. After four weeks of exposure to lunar material, 230 mice and 120 quail were autopsied by another team. Like the ‘‘lower animals,’’ the inocu­lated mice were found to be normal, and the quail that had lunar soil mixed in their feed showed no adverse reactions. The reports from these teams were greeted with a sigh of relief from all 142 sample PIs and the scores of coin­vestigators waiting anxiously to receive their allocated portion of the returned samples and get on with their analyses. If some pathogen had been found, we might still be waiting to study Apollo 11 ’s lunar treasure. The samples were declared safe for distribution around the world and were released on Septem­ber 12, 1969.

The time between the conclusion of the contamination tests on August 22, 1969, and the release of the samples twenty-one days later was spent in prelimi­nary analyses and preparing the specific types of samples required by each of the sample PIs. On August 27 the Lunar Sample Analysis Planning Team (LSAPT), chaired by Gene Simmons, issued a final internal ‘‘summary report’’ on its findings from the study of a small selection of the samples.13 This team, consisting of scientists with differing backgrounds from MSC, USGS, and other government and university laboratories, was the first group to examine lunar samples before they were released to the sample PIs.

This final summary report and the four preceding reports were read with great interest by all of us at NASA headquarters. Each report contained some new and exciting revelation. LSAPT identified two types of rocks, crystalline and aggregates (later classified as breccias), as well as a variety of fine material from the lunar soil. Although the minerals in the rocks were similar to minerals found in the Earth’s crust, there was a major difference. They contained a larger percentage of refractory elements such as titanium and zirconium. To a miner­alogist this finding was important, leading LSAPT to proclaim that this mineral assemblage provided “difficulties for the fission hypothesis,” that is, that the

Moon had been torn away from an early Earth by some cataclysmic event. If this had occurred, the minerals found in the samples should have been similar to those found on Earth.

Another, less hypothetical, conclusion was that the crystalline rocks were basalts, yet their density was greater than the average density of the Moon as a whole. This finding made it difficult to conclude that the Moon was a differenti­ated planetary body like the Earth, as it was thought to be, where the heavier material would be expected in the interior and the overall density of the planet should be higher than the density of rocks found at the surface. But this finding was consistent with the discovery of the “mascons,” since this dense material was found in mare basins. If one pursued this logic, then some large portion of the Moon must be made up of less dense material to account for the difference, or else the Moon’s core, if it had one, would have to be very small. It seemed clear that at some point part of the lunar surface had been molten.

After LSAPT performed its functions, it combined forces with the Lunar Sample Preliminary Examination Team to do more complete analyses and publish the results. To some degree this report skimmed the cream from the discoveries that would be announced later, but it served the important function of preparing us for the next missions. If we had had to wait for the sample PIs to report their findings we would have had little chance to modify or change the experiments and sampling procedures for Apollo 12 and the later missions. The LSPET report, published in Science two months after Apollo 11 returned, listed eighteen conclusions.14 The most important from my perspective, paraphrasing the report’s language, were that the crystalline rocks were different from any terrestrial rock and from meteorites; that the absence of hydrated minerals indicated there had been no surface water at Tranquility Base at any time since the rocks were exposed; that radioactive age dating showed they were crystal­lized 3 to 4 billion years ago; and that there was no evidence of biological material in the samples. Additional details and new findings would be released by the sample PIs four months later.

The Apollo 11 Lunar Science Conference was held in early January 1970 at the Rice Hotel in downtown Houston. The conference was an exciting time for all of us who had helped develop the Apollo science program. Apollo 12 had returned to Earth just a little more than a month earlier, but all of its samples were still in quarantine and unstudied. Only the Apollo 11 samples had under­

gone detailed examination by January 1970. Gary Latham, the principal inves­tigator for the passive seismic experiment, had published a short report on his findings by this date along with the LSPET report mentioned above, but the sample PIs had agreed to withhold their findings until this meeting. Those performing the detailed sample analyses were all gathering at the same place for the first time. Approximately 1,100 PIs and their collaborators, including teams from sixteen foreign countries, had spent the past three months working fever­ishly to have their analyses ready for this day.

The expectation was palpable the first morning as we milled around in the hotel lobby. Whose theories would be confirmed, whose relegated to the dustbin of lunar science? Would any of the LSPET findings be challenged or changed?

Gene Shoemaker, representing his team from USGS and several universities, made the first presentation. He described the geologic setting of the lunar samples collected by the Apollo 11 astronauts, coining the term ‘‘lunar regolith’’ for the surface characteristics at the landing site. The upper, regolith layer had been constantly churned and pulverized by impacts of all sizes. All the material returned had been collected from this fragmental debris layer, and the astro­nauts’ observations had been made within 125 feet of the landing site. No ‘‘bedrock,’’ or material in place, had been sampled. By geological standards it was not a very good collection of samples for such a large body as the Moon, but the consensus was that the samples were representative of a much larger area because of the mixing and transport of material brought in from afar as impact ejecta. Finally, he described the efforts to fix the location of each sample station. This had not been completely successful because the time limits for the EVA had restricted the number of photographs taken, but most had been located. Of the forty-seven pounds of material returned, approximately fifteen pounds had been distributed for analysis. (For the formal proceedings of the conference Shoemaker’s presentation was modified and published as ‘‘The Apollo 11 Sam­ples: Introduction.’’)15

Four days and 180 papers later the conference ended. We now had the first comprehensive view of one spot on the Moon based on data collected on the Moon itself. Several new minerals had been found, lunar lavas and breccias were common, and many samples bore evidence of shock metamorphism caused by impacts. Science devoted its entire January 30, 1970, issue to the conference. Though it is four times the size of a normal issue, it is a much more compact reference than the three-volume Proceedings for those who want to review the results of the first analyses of the Apollo 11 samples in some detail.

The oldest samples dated gave radiogenic ages of approximately 4.7 to 4.9 billion years B. p. (before the present). Others gave dates of 4.13 to 4.22 and 3.78 billion years (some of the older dates were later disputed), in general much older than the first dates offered by LSPET. Only traces of carbon were found (one anomalous sample contained almost five hundred parts per million), and there was no evidence of any bio-organic compounds. One group of investiga­tors (R. D. Johnson and C. C. Davis) stated that some of the high carbon readings might be attributable to contamination introduced during sample preparation or to errors in analytical techniques.16 They suggested that an upper limit of ten parts per million would be correct for indigenous lunar organic material. They thought the small amounts of carbon detected in some of the samples might have come from the solar wind or from carbonaceous chondrites that had struck the Moon in ages past.

Water was not identified in any of the minerals analyzed, nor did Luis Alvarez find any magnetic monopoles. Some samples studied for remnant mag­netism seemed to indicate that the Moon once had a small magnetic field, perhaps 1,000 to 1,500 gammas, or about one-thirtieth of the current field of the Earth. The present magnetic field was much smaller, however, on the order of 10 to 30 gammas, the latter figure coming from the magnetometer at the Apollo 12 site that returned data by the time of the conference.

Preliminary results from measurements of the Laser Ranging Retro- Reflector were also reported. Accuracy in measuring the Earth-Moon distance had improved over that included in Mueller’s report to the president four months earlier. This distance was now known to a precision of approximately one foot and was predicted to improve shortly to about six inches.

The Solar Wind Composition experiment carried on Apollo 11 was not discussed at the conference. This experiment, mentioned in chapter 7, consisted of a sheet of aluminum foil hanging from a pole. After being exposed for seventy-seven minutes on the lunar surface, it was retrieved and brought back to Earth and placed in quarantine in case some lunar soil had adhered to the foil. When released from quarantine, it was carefully packed and sent to Swit­zerland for analysis by its PI, Johannes Geiss. He had made a quick analysis of the gases captured on the foil, finding noble gas ions as expected, and had reported his results in December in Science.17 Eventually he extended his Apollo 11 findings based on data returned from the next four missions, examination of pieces of Surveyor 3 returned by Apollo 12, and data from the Vela satellites. Compiling all this information after his last experiment returned from the Moon, he stated in 1972 that he was now able to make good approximations of the average solar wind-noble gas abundances and ratios.18 He forecast that a better understanding would evolve of the abundances of noble gases in the Sun and the atmospheres of Venus, Mars, and the major planets.

Latham’s passive seismic experiment included in the Early Apollo Scientific Experiments Package continued to operate intermittently for twenty-one days. It survived the first lunar night before succumbing to the heat of the second day. Initial data telemetered to Earth had caused some consternation in Latham and the other members of his team. The Moon, based on these early data, seemed to be highly active seismically (apparently recording many small moonquakes), contrary to what had been predicted. After the first data had been analyzed, Ed Davin remembers walking between the Mission Control Center and the press conference room at MSC with Frank Press and Maurice Ewing, two of Latham’s coinvestigators. They were trying to figure out what to tell the assembled re­porters about this unexpectedly active Moon, apparently more active than the Earth. They asked Ed for his opinion, and he recalls being shocked that two of the world’s leading seismic authorities would ask a lowly civil servant such a profound question. Ed could not suggest a solution, so Press and Ewing ended up announcing that the Moon appeared to be more active than the Earth, a new and disturbing “scientific discovery.’’

Eventually the explanation for this totally unexpected finding became clear. The lunar module landing stage, left behind when the astronauts departed, was creaking and groaning under the thermal stress of the wide temperature swings between lunar day and lunar night. In addition, the LM and backpacks dis­carded on the surface continued to emit gas long after the astronauts departed. Each quiver and burp of gas was being detected by the extremely sensitive seismometer just sixty feet away. These disturbances appeared in the data stream as small moonquakes. No one had anticipated that such tiny movements would be measured. Thus does science advance as we try to fit new data into old theories: some mysteries are quickly resolved.

The Moon, in fact, is seismically quiet (as opposed to Earth, where large or small earthquakes are being recorded almost constantly), and this was shown again and again as we deployed four more seismometers. Once the residual effects of the SIVB and LM impacts that occurred on later missions and the astronauts’ presence had faded, the Moon stopped shaking. It was disturbed frequently by small movements believed to be caused by lunar tides (move­ments in the Moon’s crust as a result of Earth-Moon interactions), thermal changes at sunrise and sunset, small impacts, or what were interpreted as rockfalls on nearby crater rims or mountainsides. A few larger true moon – quakes were also recorded, with widely scattered epicenters concentrated at a depth of five hundred to six hundred miles, believed to be the base of the lunar mantle. The man-made shocks from the SIVB impacts also contributed to determining the thickness of the lunar crust.

Based on several years of data analysis, Latham and his team drew a number of conclusions. Below the thick lunar crust and mantle, constituting a “dynam­ically inactive outer shell,’’ was a ‘‘core’’ with ‘‘markedly different elastic proper­ties,” and the core was very small. They believed that the core was at or near the melting point, but this did not ‘‘imply a major structural or compositional discontinuity as it does for Earth. However, the presence of a true core. . . is not precluded by present data.’’ They also believed that ‘‘the presence of a thick lunar crust suggests early, intense heating of the outer shell of the Moon.’’19 This last conclusion seemed to be validated by the visual evidence of widespread maria that filled all the low elevations on the Moon’s nearside. Recent results from the Lunar Prospector mission appear to confirm Latham’s findings and indicate that the Moon’s core probably contains less than 4 percent of its mass, whereas the Earth’s core makes up 30 percent of its total mass.

Continuing now from the findings above, where do we stand in answering the questions that had perplexed many noted scientists for centuries? Most students of the Moon would agree, I believe, that satisfactory answers are now in hand for most of those questions, although there is still no unanimous interpretation. Why should study of the Moon be different from other scientific controversies?

The burning question before the unmanned and manned missions—whether the craters observed were mostly of impact or volcanic origin—had been re­solved to the satisfaction of most lunar students long before the first Apollo landing. Impacts were the answer, and Apollo data confirmed this conclusion. But the returned samples clearly showed that lava sheets or flows covered large areas of the Moon. What mechanism caused these flows is a little more debata­ble. Heating and melting of the lunar crust and mantle as a result of huge impacts is the favored explanation, not volcanic eruptions.

Next, where did the Moon come from? There is still some debate on this, but the possibilities have been narrowed and a preponderance of opinion favors one origin. Lunar samples show that the Moon’s composition is similar to that of Earth, yet different. The Moon is not compositionally exotic, as proposed by Harold Urey and others, thus it probably was not captured early in Earth’s history after having been formed somewhere else in the solar system. That leaves two theories: that it formed separately at about the same time as the Earth or that the Moon was split off from Earth by some event early in the Earth’s formation.

Because the mineral assemblages found in lunar samples differ somewhat from rocks that have formed on Earth, either origin is possible. However, the Moon most likely was torn from the Earth by the impact of another large body that contributed some of its material to the Moon, thus accounting for the mineralogical differences. This latter theory is gaining more and more favor in recent years as other conditions, such as the Moon’s angular momentum, be­come better understood and are factored into the models being used.

The next question, How old is the Moon? can now be answered with some certainty. Age dating of lunar samples has shown extremely old ages, some as high as 4.4 to 4.5 billion years B. p. This rivals the oldest ages found in mete­orites, which until this point were the most ancient objects dated. This date agrees with the thinking of most solar system students about when the solar nebula began to clump and form the planets, indicating that the Moon formed almost simultaneously with the Earth at a very early point in the birth of our solar system. The ‘‘genesis rock,’’ collected on Apollo 15, is almost pure anortho­site, a type of rock formed on Earth at great depths. It is believed to represent a piece of the Moon’s early crust. Argon-argon dating found an age of crystalliza­tion of approximately 4.0 billion years b. p.20 However, this type of dating can produce lower than actual ages; thus the ‘‘genesis rock’’ may be older-closer to 4.4 to 4.5 billion years.

Whether there has ever been water on the Moon, or whether water still exists there, has been a continuing and intriguing question. None of the samples analyzed showed that water was present during the formation of the lunar crust. But in March 1971 John Freeman of Rice University, the PI for the Suprathermal Ion Detector Experiment (SIDE), reported that he had recorded

the occurrence of water vapor for three ‘‘events’’ at his instruments left at the Apollo 12 and Apollo 14 sites. These measurements had been made at the same time Gary Latham recorded a swarm of moonquakes, suggesting that the two events were connected. Earlier Freeman had recorded the LM and SIVB impacts as disturbances in the Moon’s ionosphere, but these events had a different character than those he believed indicated water vapor. This created a stir in the media that prompted us to try to put Freeman’s claim in a larger context.21

Acknowledging the importance of discovering water on the Moon, we dis­cussed potential sources of the inferred water vapor, possibly related to material left behind by the astronauts in the LM descent stage tanks, portable life – support system tanks, and other items discarded on the lunar surface. We also pointed out that the SIDE experiment identified the mass of ions (in a gas cloud) only in a range of energy that would also include methane or neon, which could also have a lunar origin. Ultimately Freeman’s recordings were not considered conclusive in detecting water.

The recent lunar probe, Lunar Prospector, appears to support the possibility that water, in the form of ice, exists on the Moon in the permanently shadowed craters near the poles. If ice is present, it is most probably a by-product of comet impacts. Sensors on Lunar Prospector detected hydrogen, and the most likely source of the hydrogen is considered to be ice. Perhaps Freeman had detected an early whiff of water vapor from his two experiments.

To sum up the operational accomplishments of the six Apollo landing mis­sions: almost 5,000 pounds of experimental equipment were landed on the Moon, and 840 pounds of lunar material (rocks, dirt, drill cores, etc.) were returned under carefully controlled conditions. Five ALSEPs, which included most of the total of fifty-three individual experiments deployed by the astro­nauts while on the lunar surface, were placed at different locations. And ap­proximately sixty miles of traverses were recorded, both on foot and using the LRV, in support of the field geology studies and geophysical surveys. In addi­tion, detailed data were collected on missions 15, 16, and 17 from instruments carried in the command and service module, including photographs, composi­tional analysis of broad areas of the Moon’s surface, mapping its magnetic and gravity fields, and analyzing its tenuous atmosphere. All of these data contrib­uted toward deciphering the Moon’s many mysteries as well as resolving less controversial issues.

For young engineers dreaming of one day building lunar bases, the Moon will be a friendly place. Lunar bulldozers and backhoes will be able to excavate and move lunar soil just as we move soil on Earth. There will be obvious differences, but we gained sufficient data through the soil mechanics experi­ment and other experiments to design such machines. Structures could be covered with lunar soil to shield them from solar flares and high energy parti­cles, thus obviating the need to bring shielding from Earth. If needed, ‘‘regolith blocks’’ could be made from the soil that would be as useful as terrestrial cinder blocks. Unlike bases built in Antarctica, the closest terrestrial analogue to lunar bases, which must be constantly refurbished or rebuilt because of damage from snow and ice, lunar bases once constructed should last for the ages. Only a direct hit or near miss from a meteorite could damage the base. And perhaps if bases are built near the Moon’s poles the Moon can be mined for water, the most valuable of all lunar resources. The Apollo program provided the shoul­ders to stand on—now it is up to future explorers to go beyond our ‘‘giant leap for mankind.’’

A few more words concerning the results of the Clementine and Lunar Prospector programs. Both of these programs continue to add to our knowl­edge of the Moon. In some instances they are expanding on what we learned from Apollo, and in other exciting ways they are providing new information. Rather than my attempting to summarize their results to date, references in the notes discuss some of the findings.22 Many other papers and reports discuss the results of these two missions.

The final maneuver for Lunar Prospector, a last-minute addition to its scien­tific objectives, was a controlled crash similar to those carried out by the Apollo LM ascent stages and SIVBs. This time the impact point selected was a per­petually shadowed crater near the Moon’s south pole, in the hopes that tele­scopes in orbit or on Earth would record the plume from the crash and confirm the presence of water. Such a cloud was not seen, repeating our experience during the Apollo missions when I asked observatories in France with large telescopes to try to observe and measure the impact of the Apollo 16 SIVB stage. This would have been a much larger event than was expected for the Lunar Prospector impact. The time of the Apollo 16 SIVB impact prevented any United States observatories from participating, since the Moon would be below the horizon. The weather was not completely cooperative when observatories at Meudon, Pic-du-Midi, and Nice attempted to observe the impact on the night of April 19, 1972, and this might have accounted for the negative report we received.23 However, the failure to see a cloud at the impact point selected for Lunar Prospector’s final act will not detract from its successes; further analysis of data recorded by the spacecraft’s sensors will without doubt continue to add to our understanding of the Moon in the years ahead.

In successfully undertaking the challenge set by President Kennedy (with emphasis on “successfully”), Apollo taught us one final lesson. Apollo’s heritage went far beyond knowledge about the Moon and Earth. Now that many of the records of the former Soviet Union have been opened to public scrutiny, it has been confirmed that we really were in a race to the Moon. It certainly seemed that way to us at the time, but you could not be sure because Soviet launches were always veiled in secrecy; the world became aware of them only after they were on their way to whatever destination, and failures were never reported. The Soviets’ long-range plans were seldom discussed, although Boris Voishol, from the Soviet Tectonic Academy, writing in the September 1968 Geotimes, stated: ‘‘The first landing of Soviet cosmonauts on our moon is scheduled in the near future.’’24

Based on information available at that time, the missing ingredient in their ability to send men to the Moon was a booster as large as the Saturn V, which would be needed for the round trip. Without such a rocket we assumed that if they were really intent on a manned lunar landing they would use their smaller, proven rockets to assemble the needed launch capability in Earth orbit before going on to the Moon-one of NASA’s original proposals. We now know that they were building a Saturn V-class rocket but that on its first test flight it crashed shortly after lift-off. On a second launch attempt a few months later, it exploded on the pad, apparently killing some of their rocket experts, and was never rebuilt.25 The Soviet failures-and there were many-were only a matter of speculation for most of us, though undoubtedly there were some who were privy to intelligence sources and knew about their difficulties. Our launches, successful or unsuccessful, were always made in full view of the world.

What if the Soviet Union had landed men on the Moon first? Several writers have discussed the effect of Project Apollo on the Soviet Union; here is another view with which you may or may not agree. My father, a civilian stationed in West Germany for the Army Signal Corps at the time of the launch of Sputnik I, remembered an unnerving encounter with one of his German contractors. When it was confirmed that the Soviets had successfully orbited the first satel­lite, this man came running up with fear on his face. His conclusion was that this demonstration of Soviet technological superiority spelled doom for the world. The United States failures at launching the Vanguard rocket were well known. Suddenly the Soviet Union had leapfrogged our efforts. Along with its newly demonstrated nuclear weapons, this made the man believe the bad guys had won the Cold War. We would soon have to knuckle under to this new dominant world force. He was seeking reassurance that his analysis was wrong, but with limited knowledge of how our space programs were proceeding, my father could not give it.

The point of this anecdote is to show how fragile a nation’s leadership is in a rapidly evolving world. In view of their recent history, West Germans in Octo­ber 1957 might be forgiven for being pessimistic. But as I remember, this pessimism was widespread even in the United States, with finger pointing and blame all around for our inability to beat the Soviets during the early days of space flight.

What would the world look like today if the Soviets’ program had not experienced its hidden failures and they had been first to land men on the Moon? I suspect it would be different, but of course there is no way to prove it. Everyone likes a winner and gravitates toward one regardless of worthiness; second place seldom attracts much enthusiasm. Accommodation to Soviet leadership would have been rationalized, and the Soviet bloc might have be­come the dominant force in world politics, perhaps postponing or averting its ultimate economic collapse. Meanwhile, we would be scrambling to catch up and demonstrate that a democratic government could do as well as or better than a state-directed totalitarian government. Remember, in the 1960s many countries were experimenting with or embracing communist forms of govern­ment, and Soviet-led expansion of communist ideology was making great strides even without demonstrating the overall technological leadership that a ‘‘first’’ on the Moon would have given.

What is the lesson of Apollo that goes beyond being the first to land on the Moon and the expansion of our scientific knowledge? It seems pretty basic. Free societies can successfully undertake enormously complex actions—if they dare. Although the United States was the leader in Apollo, many other nations con­tributed people, technology, or facilities. Apollo was a dream that everyone could embrace, if permitted, and all could share in the sweet reward of success. The few words on the plaque carried by Apollo 11 said it all: it was an accom­plishment ‘‘for all mankind.’’

Conceived primarily as a political statement, Apollo achieved much more than its original goal. Now, when faced with seemingly intractable problems, someone will be heard to say, ‘‘If we can land a man on the Moon, why can’t we [fill in the appropriate objective]?’’ And of course that is the right question to ask, because people of goodwill, working together, are capable of solving very difficult problems. Apollo proved it. Let’s not forget that dreaming big has its own rewards, even if occasionally we stub a toe. That is the essential lesson I carry away from my Apollo days, and I hope it will be remembered by those who study and follow our example in the future.

Taking Science to the Moon

The technical achievements that permitted the National Aeronautics and Space Administration (NASA), other government agencies, and their contractors to fulfill President John F. Kennedy’s promise of ‘‘landing a man on the moon and returning him safely’’ have often been described. Most previous authors have included anecdotes that enhance our appreciation of how Project Apollo was successfully accomplished, although many are retold at second or third hand. Several movies such as The Right Stuff and Apollo 13 showed both true and fictional accounts of the spirit and engineering skills that characterized the entire project, focusing primarily on the major or well-known participants.

A story that has not been completely told, however, is how a small band of somewhat anonymous NASA staffers, allied with scientists inside and outside government, struggled to persuade the management of NASA to look beyond the initial Apollo landing and reap a scientific harvest from this historic under­taking. Here is that story as seen through the eyes of a participant based at NASA headquarters—a pack rat who kept many of the internal memos, reports, photos, and notes that document that ten-year struggle. It highlights the contri­butions of many of those who worked with me during the Apollo program. Some of them have received little public recognition for their efforts. I hope that this insider background will give readers a better understanding of the behind-the-scenes maneuvering that led to many of Project Apollo’s scientific achievements, which have enriched our understanding not only of the Moon but, more important, of the small planet we call Earth.

Acknowledgments

Many people and organizations helped and encouraged me while I was writing this book, and they deserve credit. Although I had saved many boxes of material I collected during my Apollo days at NASA in anticipation of one day writing this story, I soon found this source material was insufficient. Calling old col­leagues to ask if they had kept records was not very fruitful at first, but even­tually I was successful.

The first person who agreed to share his records covering part of this period was Robert Fudali, who was on the Bellcomm staff during Apollo’s early days. His material not only contributed to the accuracy of this story but served as a valuable reminder of some of the events that occurred during the formative years of Apollo science. I have quoted liberally from a few of Bob’s colorful internal memos.

Gordon Swann, a friend, former colleague, and principal investigator who took part in the struggle to develop science payloads for Apollo, especially those aspects related to the astronauts’ geological investigations, reviewed early drafts and provided many important comments and suggestions as well as a few of his famous anecdotes—some printable, some not. Gordon should be encouraged to one day write his account of Apollo.

Paul Lowman, who figures prominently in this story, was an invaluable source of material and a resource for clarifying many events. Paul is renowned among his NASA colleagues as a pack rat of the first degree: his office is so filled with reports and trivia that when you first enter it is hard to find his desk. However, his propensity for maintaining his archives has benefited many who have written about NASA’s early days. He also reviewed the manuscript and offered many useful comments.

James Downey, Herman Gierow, Farouk El Baz, and Charles Weatherred reviewed drafts at various stages, and Jim spent many hours going through the files at the Marshall Space Flight Center library to select material relating to the early years of our post-Apollo work. Chuck Weatherred and Eugene Zaitzeff (both Bendix employees during Apollo) and Charles Spoelhop at Eastman Kodak also provided important background material from their files. My for­mer colleagues Philip Culbertson, Richard Allenby, Edward Davin, Richard Green, George Esenwein, Alex Schwarzkopf, Saverio ‘‘Sonny’’ Morea, George Ulrich, Raymond Batson, William Muehlberger, Floyd Roberson, and John Bensko took the time to provide information and pictures and to confirm recollections now more than thirty years old. Hugh Neeson, a former Textron – Bell engineer, searched the archives of the Niagara Aerospace Museum to find rare artists’ drawings of the lunar flying vehicle. Bruce Beattie, my son, became a fact finder after I moved from Maryland, following up on questions that could be answered by Washington sources.

The NASA headquarters history office, in particular Lee Saegesser (before he retired) and Roger Launius and his staff, helped me access the records still maintained in Washington. Glen Swanson, NASA Johnson Space Center (JSC) historian, provided key contacts at JSC, including Joseph Kosmo at the Flight Crew Support Division and Judith Allton in the lunar sample curator’s office that allowed me to fill in a few blanks in my story. And most important Michael Gentry and David Sharron at the JSC Media Resource Center, who spent con­siderable time helping me select and acquire the photos and drawings in the book.

Roger Van Ghent, a colleague and fellow Floridian, advised me on the intri­cacies of using my computer to ease my writing load and also helped compile the index.

To all these people and the many colleagues and friends whose names do not appear, my sincere thanks for your help and encouragement from my first days at NASA until the present.

Finally, I thank Alice Bennett at the University of Chicago for editing and improving the manuscript and Bob Brugger, my editor at the Johns Hopkins University Press, for running interference and patiently guiding me through the publishing process. There is no substitute for an unflappable editor.

Introduction

Anchored to its launch pad on the morning of July 16, 1969, and scheduled to launch Apollo 11 on our first attempt to land men on the Moon, the fully fueled Saturn V launch vehicle weighed over six million pounds. From the nozzles at the base of the giant S-1C first stage to the top of the solid rocket-propelled escape tower, it measured 363 feet. In 1962, one year after President Kennedy had given the go-ahead for Project Apollo, the critical decisions had been made on how to execute his difficult challenge. Saturn V, with its multiple stages, was the key to reaching the goal, the product of seven years of effort by hundreds of thousands of government and contract workers.

The original planning in 1960 and 1961 centered on building a huge rocket to launch a spacecraft directly from Earth to the lunar surface, followed by a direct return home. The mission design finally selected was very different. It required a smaller, but still very large, multistage rocket to launch three astro­nauts into a low Earth orbit and then send them on to the Moon in a spacecraft that combined command and logistics modules with a lunar lander. On arriv­ing at the Moon, these combined spacecraft would be parked in a low lunar orbit. The lunar lander, a two-stage (descent and ascent stages) two-man space­craft, would then separate and go to the lunar surface. The command and service module, with the third astronaut on board, would remain in lunar orbit to rendezvous and link up with the astronauts when they returned from the Moon’s surface. After the astronauts who had landed on the Moon transferred back to the command module, they would jettison the lunar lander ascent stage, and all three would leave lunar orbit and return to Earth in the command module for an ocean recovery.

Lunar orbit rendezvous (LOR) was the unique feature of the mission design

that allowed NASA to reduce the size of the initial launch vehicle. An LOR flight profile required the development of a new, powerful rocket (Saturn У) and the design and fabrication of two complex spacecraft that would perform a series of difficult and potentially dangerous space maneuvers never before attempted. But a manned lunar landing designed around LOR was sold to NASA manage­ment as the quickest, least risky, and lowest-cost way to carry out the president’s mandate. The LOR decision fixed the broad architecture of the mission and defined the parameters within which the scientific community would have to work when NASA finally determined what scientific activities were appropriate for future Apollo astronauts to carry out. (How NASA decided to adopt LOR, in a behind-the-scenes debate, has been covered in some detail in several of the references cited.)

Because the president’s mandate did not require that any specific tasks be accomplished once the astronauts arrived on the Moon, the initial spacecraft design did not include weight or storage allowances for scientific payloads. Somewhere, somehow, amid the six million pounds and 363 feet, we would have to squeeze in a science payload. The earliest thinking was, ‘‘We’ll land, take a few photographs, pick up a few rocks, and take off as soon as possible.’’ The need to do much more was not considered in the planning. For many NASA engineers and managers the lunar landing was a one-shot affair. After the first successful landing, NASA would pack up its rockets and do something else. Why take any more chances with the astronauts’ lives on this risky adventure? This thinking was soon to change, at least in some circles.

The first officially sanctioned attempt to change this thinking took place in March 1962 when Charles P. Sonett, of the NASA Ames Research Center in California, was asked to convene a small group of scientists to recommend a list of experiments to be undertaken once the astronauts landed on the Moon. This meeting, requested by NASA’s Office of Manned Space Flight, was held in conjunction with a National Academy of Sciences Space Science Board Summer Study taking place at Iowa State University in Ames so that the Academy’s participants could review and comment on the recommendations Sonett’s team would make. The Sonett Report, submitted to NASA management in July 1962, became the foundation for all subsequent lunar science studies and recommen­dations. Circulated in draft form at NASA and other organizations throughout the rest of 1962 and most of 1963, the report elicited both support and crit­icism. It is at this point in the evolution of Apollo science, with a short digres­sion to set the stage, that I became involved, and here I take up the story.

Each chapter is written as a somewhat complete account of its subject. The chronology for a given chapter is correct as events unfolded, but there is some overlap in time as we move from one chapter to the next. I hope this will not be confusing but will provide a better perspective on how the individual pieces of the lunar science puzzle came together. I have also attempted to explain the roles of the key contractors and give credit to some who worked with us from the very beginning as we struggled to define and build the many experiments and supporting equipment that eventually made up the Apollo science pay­loads. I believe that most accounts of the Apollo program fail to give enough recognition to the many contractors who were essential contributors to the project’s success.

One of the major players in this story was the late Eugene M. Shoemaker. Gene was involved in almost every aspect of Apollo science and had graciously agreed to review this manuscript when it was ready. I was greatly anticipating the comments and critique of this friend and colleague, hoping he could refresh my memory and suggest additions or changes for accuracy. But before I could send him an early manuscript, Gene died tragically in an auto accident on July 18, 1997, while studying impact craters in Australia. He will be fondly remembered and greatly missed. Not only was he an outstanding scientist who shaped our thinking on many subjects, including how we should explore the Moon, he was also a brilliant teacher whose greatest legacy, perhaps, will be the many young (and old) scientists and engineers who will follow in his footsteps and lead us back to the Moon and beyond—to Mars and the far reaches of our solar system.

Taking Science to the Moon

From the Jungle to Washington

In February 1962 John Glenn was at Cape Canaveral preparing for his attempt to become the first American to orbit the Earth during the Mercury program. I was working for the Mobil Oil Corporation as an exploration geologist super­vising a small field party in the rain forest of northern Colombia. Even in this remote area I could pick up Armed Forces Radio and the Voice of America on my battery-operated Zenith Transoceanic radio and stay up to date on the major events of the day. We had been closely following the launches of the newly formed National Aeronautics and Space Administration, and along with everyone back in the United States, we were disappointed at the failures and delays as we tried to catch up with the Soviet Union’s aggressive space program.

After each of the several launch delays for Glenn’s flight, NASA would project a new liftoff time, and based on these projections we would try to complete our daily fieldwork and get back to camp to hear the launch broad­cast. Far from home, with our immediate world bounded by a small rain forest camp and how far we could ride each day on the back of a mule, it was easy to become absorbed in the drama at Cape Canaveral. One day, during one of the several holds before Glenn’s launch, the announcer filled some airtime by interviewing someone from NASA’s Public Affairs Office. During the interview Project Apollo was discussed (what little was known of it at the time), and it was mentioned that for the Moon landings NASA would need to hire geologists to help plan the missions. He gave an address where those interested could apply. My curiosity was piqued. I copied down the address, pulled out the rusty typewriter we used to write our monthly reports, and composed a letter to

NASA. I explained that I was not only a geologist but a former navy jet pilot and said I thought I would fit right in with NASA and all the astronauts.

Eventually John Glenn was launched successfully. When I next went to Bogota I mailed my letter, convinced that NASA could not turn down such outstanding qualifications. In my naivete I thought I might even have a chance to become an astronaut. Who had a better combination of experience to go to the Moon, I reasoned, than a geologist-jet pilot, especially one accustomed to working in strange places under difficult conditions (coexisting with army ants, vampire bats, and jaguars)? With some modesty, my letter implied this interest. It was several months before I had a reply from NASA—a polite letter thanking me for my interest. To be considered, I must fill out the enclosed forms and submit my application to the Goddard Space Flight Center in Greenbelt, Mary­land. I did so, and the wait began—with some anticipation, given NASA’s encouraging reply.

With the start of the rainy season I was back in Bogota when another envelope arrived telling me I had qualified as a GS-13, aerospace technologist- lunar and planetary studies, and that my application was being circulated within NASA to determine if a position was available. I wasn’t sure what an aerospace technologist was, but it sounded impressive. I had visions of being asked to do exciting things at this new agency with the improbable task of sending men to the Moon. Then began a longer wait. In December I received another letter saying that no positions were open but that they would keep my application on record in case one turned up. Rejection! That didn’t fit in with my plans, and I resolved to pursue my quest the next time I was in the United States.

My next leave came in June 1963, and I decided to go to Washington to talk directly to someone at NASA. I bought an aerospace trade journal listing the latest NASA organization, complete with names. In it I found an office at NASA headquarters that sounded as if my background and interests would fit—Lunar and Planetary Programs in the Office of Space Science, headed by Urner Lid­dell. From my family’s home in New Jersey I drove to Washington and, without an appointment, went to Liddell’s office. He was traveling that day, but his deputy, Richard Allenby, was in. This was great good fortune, since Liddell turned out to be a rather formal bureaucrat who probably would not have seen me without an appointment. Dick Allenby was just the opposite and agreed to interview me. After briefly introducing myself, I learned he was an old oil field hand (geophysicist) who had worked in Colombia just a few years earlier, and we had several friends in common. We hit it off at once, marking the beginning of a long professional and personal relationship. Dick liked my background but had no openings. He then set up a meeting with navy captain Lee Scherer (another former pilot), who had just been hired to manage the Lunar Orbiter program (satellites that would orbit the Moon to photograph potential Apollo landing sites). He also was not hiring at the time, but he thought someone in the Office of Manned Space Flight needed a person with my experience. I was beginning to question my timing: lots was going on at NASA, with new offices being set up all over town, but just as the last NASA letter stated, no one had an opening. Lee, who would become my boss six years later, set up a meeting with another military man newly detailed to NASA, Maj. Thomas C. Evans, U. S. Army Corps of Engineers.

Tom Evans was an impressive officer, later to become a congressman from Iowa. Tom had been the officer in charge of establishing Camp Century in Greenland, the first successful adaptation of nuclear power for a military ground base. His background was ideal for his job at NASA-designing a future lunar base. After Lee Scherer’s introduction got me in the door, he spent the next hour or so telling me about his new office’s responsibility—planning a lunar program to follow a successful Apollo program. He was enthusiastic and brimming with ideas, the kind of leader everyone looks forward to working for. Best of all, he thought I could help the team he was putting together. Since it was getting late in the day, Tom asked me to return the next morning to talk to his deputy, Capt. Edward P. Andrews, U. S. Army, and determine how we could proceed.

My discussion with Ed Andrews went well, and since I had already received a civil service job rating, he proposed starting the paperwork to hire me. Two days in Washington and I was being offered a job as a lunar aerospace technolo­gist at what I considered the most exciting place in town! It would mean a pay cut from my Mobil salary (I would receive the princely sum of $11,150 a year), but I couldn’t pass up the opportunity. Ed took my paperwork and told me he would call me in Colombia when everything was final; he didn’t see any reason the position would not be approved and said I should plan on moving my family to Washington.

Returning to Colombia in July, I took Ed at his word and began to close out my work. My supervisor knew about my plans, of course, since I had listed him as a reference. My coworkers all thought I was crazy to take on such a job; most thought trying to get a man to the Moon was quixotic at best and probably impossible. Planning what to do after we landed on the Moon was real science fiction. I thought they were all being short-sighted and that they would be missing out on the beginning of a real adventure. In August I got the phone call I was waiting for. Ed Andrews said all was in order and they were waiting for me to arrive. With a smug smile I filed away my NASA correspondence, including the rejection letter, and at the end of August my family and I left Colombia to begin a new calling-one that never lost its thrill and satisfaction over the next ten eventful years.

And so I began my career at NASA; a GS-13 aerospace technologist in the Office of Manned Space Flight, Manned Lunar Missions Studies. When I ar­rived in Washington, NASA offices were spread all over town awaiting the construction of a new government building dedicated to NASA, in southwest Washington. In September 1963 our offices were at 1815 H Street NW, just a few blocks from the White House. We shared the building with other orga­nizations and other NASA offices, including program offices for manned plane­tary missions, systems engineering, launch vehicle studies, and other advanced studies.

I was assigned an office with another recent hire, Thomas Albert, a mechan­ical and nuclear engineer who was determining how to modify the planned Apollo systems to enable longer staytimes and lunar base missions. Since I came from a work environment where we primarily wrote reports based on work we had accomplished in the field or laboratory, Tom really impressed me. He would spend hours on the phone talking to NASA and private company engi­neers, taking a few notes and going on to his next call, all the while speaking a language I didn’t understand, in which every third word seemed to be an acronym. I thought I’d never understand NASA-speak, in which acronyms were the order of the day. It was annoying at first, but soon I started to catch on and quickly moved to the next level where I invented my own program acronyms. This new skill brought a real sense of control. I am convinced that NASA could not have functioned without these shortcuts, and it became an unspecified requirement that new programs come up with catchy acronyms, most pro­nounced like real words, that would appeal to the ears and eyes of management, Congress, and the media. (You’ll soon become accustomed to them as well and will have less need to consult the list of abbreviations in the front of the book.)

Our office at this time consisted of eight engineers with diverse backgrounds plus two secretaries. Except for Tom and Ed, we all shared the services of one secretary. Two or three engineers occupied each office space: new arrivals were assigned interior offices; offices with windows were for senior staff. Accom­modations were spartan, but there were few complaints since we would soon be moving to a new building. There was one empty desk in the office I shared with Tom; it had been occupied part time by Eugene Shoemaker, detailed from the United States Geological Survey (USGS), who was on his way to Flagstaff, Arizona, to start a new USGS office. I missed meeting him by a few days, but our paths would soon cross, and we would work closely together until the end of Apollo.

My first days at NASA involved the usual getting acquainted. Although during my navy service I had been a part of another government bureaucracy, NASA functioned quite differently. Owing in part to Tom Evans’s style and NASA’s being a new agency with an unprecedented mission, multitudinous rules and procedures had not yet been instituted, and the staff was given great freedom of action. Since for the past six years I had usually made my own daily schedule, this was an ideal situation for me. With Ed Andrews’s guidance I immediately began to define my role and make the contacts at NASA and in the scientific community that would make my job easier.

I soon learned that Gene Shoemaker had come to NASA to help bridge the wide gap between the science side of NASA, represented by the Office of Space Science (OSS), where I had made my first NASA contact, and the Office of Manned Space Flight (OMSF). Major differences had surfaced between OSS and OMSF over how to apportion NASA’s overall budget. The debate on how to accomplish science on Apollo still lay ahead. OMSF was already receiving the major portion of NASA’s budget, and OSS staff, as well as scientists outside NASA who looked to OSS to fund their pet projects, were constantly fighting to persuade top management to change NASA’s funding priorities. These efforts were led by such luminaries as James Van Allen, who had made one of the first space-based science discoveries—the radiation belts surrounding Earth that were later named after him. The complaints were reinforced by the National Academy of Sciences and its Space Science Board, which provided advice to Homer Newell, the OSS administrator. I was told that Shoemaker, during his brief stay at NASA, had begun to reduce some of the distrust that had devel­oped but had only scratched the surface. Apparently it would take more than his talents to resolve these differences. Despite many compromises and much cooperation, forty years later this power struggle still rages inside and out­side NASA.

Into this controversial arena I ventured and, with Tom Evans’s blessing, was given an unofficial second hat to work with both OSS and OMSF on matters dealing with lunar exploration. When Shoemaker left, Verne C. Fryklund, who had been working on Newell’s staff, took his place. Fryklund was definitely from the old school. Gruff, with a bushy mustache and a half-smoked but unlit cigar perpetually in his mouth, he usually looked professorial in a tweed jacket with leather elbow patches. Being detailed from USGS, he was given the title of acting director, Manned Space Sciences Division, Office of Space Science. His primary duty was the same as Shoemaker’s—to be the go-between for the Office of Space Science and the Office of Manned Space Flight. During his shuttle diplomacy, he was to present the interests of the science community to NASA’s manned space side, which was not viewed as friendly to science. Fryklund became my unofficial second boss. By Washington standards his title was not imposing, especially with the ‘‘acting’’ designation. His staff was appropriately small, consisting of several headquarters staffers and a number of detailees, in­cluding geologist Paul Lowman from the Goddard Space Flight Center (GSFC) and several others from the Jet Propulsion Laboratory (JPL). Thus he was receptive to having me join his office.

Fryklund, an experienced bureaucrat, approached his new job cautiously. The complicated politics were self-evident to someone with his background, and he was fully aware of the gulf between the two organizations. Until this time nothing had been officially decided about what science projects would be car­ried out on the Apollo missions. This became his first priority. Shuttling back and forth between high-level meetings at OSS and OMSF, Fryklund relied on a draft report on the scientific aspects of the Apollo program (commonly re­ferred to as the Sonett Report after its chairman, Charles P. Sonett of the NASA Ames Research Center).1 It served as his guide and point of departure to lend weight to his arguments on what needed to be done for Apollo science.

Sonett’s ad hoc working group had convened at Iowa State University in the spring of 1962 at the request of the Office of Manned Space Flight to recom­mend what scientific activities should be included on the Apollo missions. The group had twenty members and consultants with diverse scientific back­grounds, including strong representation from USGS led by Gene Shoemaker.

Paul Lowman served on the geophysics (solid body) subgroup and also helped compile the final report, while Fryklund worked with the geology and geo­chemistry subgroup during their meetings.

William Lee, assistant director for human factors in the Office of Manned Space Flight, provided guidelines at the start of the working group’s delibera­tions. These guidelines defined the parameters within which the working group would operate. They were relatively short and simple (two and a half pages), since at that time little was known about the constraints the astronauts would be operating under and since all the Apollo hardware was in an early design phase.

The working group was asked to consider experiments and tasks that could be accomplished on the Moon in periods of one hour, eight hours, twenty-four hours, and seven days. Because NASA still was not sure what the flight profiles would be, no guidance was given for any operations on the way to the Moon or in lunar orbit. Choosing landing site(s) was also not part of the working group’s charter, although its recommendations could influence site selection. Advice on power and communication capabilities for transmitting scientific data was very general, and the committee members were told that this should not restrict them. They were to plan for more than one but fewer than ten missions with the possibility of carrying one hundred to two hundred pounds of scientific pay­load. Life-support supplies would limit the crew’s operations to a radius of approximately half a mile. They were cautioned that the astronauts’ space suits might hinder their ability to perform ‘‘precise manipulations.” And finally, they were told that it might be possible to include a ‘‘professional scientist’’ in the crew, but that this would ‘‘significantly complicate our selection and training program, and [such a recommendation] should not be made unnecessarily.”

Today, reading between the lines and looking at the numbers the committee was given to work with, it seems clear that these guidelines sent a message to the members that scientific ventures during the Apollo missions might be tolerated but that they should not have high expectations. This message was repeated in the years ahead, much to the dismay of the scientific community.

Despite the restrictions, the draft report contained wide-ranging recom­mendations that included geological and geophysical experiments to be done on the Moon as well as experiments in surface physics, atmospheric measure­ments, and particles and fields. Bill Lee’s guidelines were to some degree ig­nored; the assembled scientists could not resist telling NASA what needed to be done. What they recommended could not be carried out with only one to two hundred pounds of payload, and they described geology traverses up to fifty miles from the landing site. They also detailed sample collection, including drill or punch core samples, and potential landing sites were suggested by Shoe­maker and by Richard E. Eggleton of USGS and Duane W. Dugan of the Ames Research Center. The report went so far as to describe what type of astronaut should be on the flights and the criteria for finding such recruits.

Since the report had been requested by OMSF and not by the science side of NASA, its recommendations carried some weight in OMSF offices. The draft had been circulated to participants at the National Academy of Sciences 1962 Iowa Summer Study, who had met at the same time as Sonett’s working group.2 Thus the Sonett Report would include the endorsement of the other side of NASA’s house (the scientists) when it was officially released. Although the Iowa Summer Study group agreed with the general conclusions of the Sonett Report, it recommended that the scope of the proposed investigations be more re­stricted than those spelled out in the report, a rather surprising recommenda­tion in light of later criticisms from the scientific community.

Based on these recommendations, and with his bosses in both OSS (Homer Newell) and OMSF (Joseph Shea) concurring, in early October 1963, one month after my arrival, Fryklund sent a memo to Robert R. Gilruth, director of the Manned Spacecraft Center (MSC) in Houston, containing the first official scientific guidelines for Project Apollo. As is the nature of guidelines, they established a broad framework for planning, but they provided no specifics on how long the astronauts would be on the Moon or how much payload weight should be allocated for science. These numbers were to come later. The eight guidelines included a listing of three functional scientific activities in decreasing order of importance: ‘‘a. Comprehensive observation of lunar phenomena; b. Collection of representative samples; and c. Emplacement of monitoring equip – ment.’’3 Assigning sample collection a number two priority is interesting since, as we will see, in later planning it became the astronauts’ first task once they were on the lunar surface. Back in Washington we began trying to flesh out the guidelines by reading between the lines of the Sonett Report and translating the recommendations to some hard numbers.

From the information we could collect, it was evident that the range of measurements and activities the Sonett committee had listed, even if reduced to follow the National Academy of Science’s recommendations, would require a science payload far exceeding the target of one to two hundred pounds. One month before Fryklund issued the guidelines, and unknown to headquarters, MSC jumped the gun and hired a contractor, Texas Instruments, to spell out Apollo experiments and measurements to be made on the lunar surface based on MSC guidelines. The report, when it was eventually issued in 1964, was dismissed as amateurish by headquarters and by members of the scientific community who had begun to focus on Apollo science. This difference of perspective signaled a clash between headquarters and the small MSC science staff over who would define Apollo science.

Adding to this mix of ideas on what science to carry out on the Moon, in early 1963 Bellcomm engineers had provided some analyses of potential Apollo and post-Apollo scientific operations. Bellcomm had been created in March 1962 by AT&T at the request of NASA administrator James Webb to provide technical support to NASA headquarters. By the time I arrived Bellcomm had grown to over 150 engineers and support staff and had already run afoul of MSC engineers, who accused the company of being a meddling tool of head­quarters-some at MSC went so far as to call the staff headquarters spies. MSC tried to exclude them from some meetings by keeping the schedules quiet so that when the meetings were announced it would be too late for the Bellcom – mers to make the trip from Washington to Houston. Another aspect of the visits that MSC found annoying was that Bellcomm required trip reports, so everyone who read them knew about what went on and about any disagree­ments with MSC’s proposals. Disagreements were frequent, and the second – guessing by Bellcomm continued throughout the program, often leading to positive changes, especially concerning the science payload. Eventually a small group of Bellcomm scientists and engineers were assigned to support Evans’s office, and they became important adjuncts to our small staff. Their support and numbers grew as Apollo science evolved.

At the end of January 1963 two Bellcomm staffers, Cabel A. Pearse and Harley W. Radin, presented a study examining the scientific advantages of having an unmanned logistic system deliver a fifteen-hundred-pound payload to the lunar surface. They concluded that the best use of such a system would be to provide ‘‘a fixed scientific laboratory equipped with a wide variety of sci­entific instrumentation.’’4 Two months later, under the leadership of Brian Howard, one of England’s ‘‘brain drain’’ expatriates, with Robert F. Fudali, Cabel A. Pearse, and Thomas Powers, Bellcomm issued a second report, The

Scientific Exploitation of the Moon.5 It provided a preliminary analysis of the type of science that might be conducted utilizing Apollo hardware to deliver a logistics payload of seven to ten thousand pounds to the lunar surface, the payload sizes being studied by Evans’s office. Although the second report does not cite the draft Sonett Report by name, the authors were surely aware of its existence because they include most of the experiments it described and it is cited in the January report. In addition, they recommended carrying out a variety of other operations and experiments including the use of roving vehi­cles and deep drilling. To my knowledge the Bellcomm reports and Lunar Logis­tic System, a ten-volume report issued by the Marshall Space Flight Center (MSFC) at the same time as the Bellcomm report, represent the first attempts to document the feasibility of using Apollo hardware for extended exploration on the lunar surface.6 These reports were my first exposure to such thinking and were among the early references on my NASA office bookshelves.

In late October 1963, returning from one of these frequent meetings, Fryk – lund rushed into the office we shared and announced, “They’ve just agreed; we have 250 pounds for science!’’ ‘‘They’’ being NASA Manned Space Flight senior management. Having been on the job only a few weeks and a latecomer to what had been a major struggle, I showed only muted enthusiasm. Based on my limited experience and initial looks at what a good science payload like that recommended in the Sonett Report would weigh, 250 pounds seemed a minor victory. A thousand pounds or more would have been better. But a victory it really was, certainly better than the one to two hundred pounds given to the Sonett working group. Once our foot was in the door, we quickly capitalized on the opportunity to define a complete payload based on this ‘‘official’’ number.

Other major changes had also been taking place in NASA. Headquarters was swiftly evolving. New organizations were being created almost weekly, and the staff was expanding rapidly. During 1963, the year I came, NASA headquarters almost doubled in size. With all these changes the headquarters phone directory was always out of date, and addenda were published every month. Brainerd Holmes, who until September had been in charge of manned space flight operations as director of the Office of Manned Space Flight, resigned and was replaced by George Mueller from Space Technology Laboratories. Mueller was given the new title of associate administrator, Office of Manned Space Flight, a third tier of top management just below administrator James Webb and his deputy, Hugh Dryden and associate administrator Robert Seamans. Homer

Newell was elevated at the same time to a similar position with the title associate administrator, Office of Space Science and Applications (OSSA). With his ap­pointment Mueller introduced a different management style to Manned Space Flight, one that would have a profound effect on Project Apollo’s future.

Toward the end of the year our office was merged with several others, and the new organization was called Advanced Manned Missions Programs. Ed­ward Z. ‘‘E. Z.’’ Gray was hired from the Boeing Company to be our leader, and we soon moved to our new offices at 600 Independence Avenue SW. In January 1964 Maj. Gen. Samuel C. Phillips was detailed from the Air Force Ballistic Systems Division to become Mueller’s deputy director for the Apollo program. Later in the year his title was upgraded to director.

In the wave of reorganization, Fryklund’s tenure as acting director was short lived. Homer Newell, in agreement with Mueller, formally established the Of­fice of Manned Space Science, reporting to both his office and Mueller’s. Willis Foster was brought in from the Department of Defense as the new full-time director, and Fryklund became Foster’s chief of lunar and planetary sciences. After some eight months working for Foster, he transferred back to Newell’s staff, and a short time later he returned to USGS to work in its military geology branch. Foster’s office, starting with an original staff of eight, grew rapidly (and now included Peter Badgley, my former thesis adviser at the Colorado School of Mines). Dick Allenby was transferred from the OSSA Lunar and Planetary Programs Office to become Foster’s deputy. Anthony Calio was brought in from the newly formed Electronics Research Center in Cambridge, Massachusetts, to provide some engineering muscle, and along with Jacob ‘‘Jack’’ Trombka he began to coordinate the planning for scientific instrumentation. Edward Chao, another USGS detailee, became the office expert on how to handle the antici­pated scientific treasure—the samples collected. Edward M. Davin, an acquain­tance of Allenby’s, was hired from Esso Research (now Exxon) in Houston in the summer of 1964 to join Allenby as the resident geophysicists, representing a scientific discipline that would increase in importance as the Apollo experi­ments were selected.

Will Foster now became my unofficial second boss, and I continued to work on developing the science payloads for Apollo flights as well as later undertak­ings. How we accomplished this for Apollo, and eventually went far beyond the initial 250-pound allocation, follows in the next chapters. But first, from a scien­tific perspective, why fight to get a science payload on Apollo in the first place?

Early Theories and Questions. about the Moon

If you have binoculars of ten power or even less, you can go out in your backyard on any clear night when the Moon is up—best perhaps at a quarter – moon phase, not a full moon—and become a lunar scientist. Brace yourself against a solid support so your hands are steady and focus on the line that separates the illuminated part of the Moon from the dark portion. Near this line the Sun casts the longest shadows, and you can see the greatest topographic detail. The technical term for this line is the lunar terminator, but you needn’t know this to start your studies. Your ten-power binoculars are about half as powerful as the telescope constructed by Galileo Galilei, who early in the seven­teenth century first began to study the Moon with more than the naked eye.

What will you see? Depending on where the line between the bright and dark portions falls on the particular night, you will probably see, just as Galileo did in 1609—to his amazement—some large and small circular craters, perhaps some mountains, and some apparently smooth areas that are known as maria, or seas. In 1963, some 350 years after Galileo made his first observations, the craters were the most controversial of all lunar features, sparking the most heated debates. What was their origin? Were they the remains of volcanoes? Were they caused by impacts like those that left similar craters on Earth? Were they the result of some combination of processes or the product of unknown forces? The lunar maria were also controversial; they were generally interpreted as lava flows. But how were they formed, and how did they spread over such a vast area? How were the mountains formed? Their very existence provoked debates about the internal structure of the Moon and its evolution.

The major, fundamental lunar questions being debated by planetary scien­tists when the Apollo program began can be quickly summarized: How old is the Moon, how was it formed, and what is its composition? Finding the answers was the driving force behind the desire to carry out a host of experiments on the Apollo missions. And a large science payload would be needed to resolve these difficult questions. The answers to some of them would come in part from samples collected on the Apollo landings, and in turn the samples would tell us a lot about the origin of the craters. If the Apollo missions landed at interesting points on the Moon and included various geophysical experiments along with geologic traverses, these mysteries might be resolved. From the answers we anticipated understanding Earth better, especially its early history. When I joined NASA in 1963 my knowledge of the Moon and of the ongoing debates was close to zero. I quickly resolved to fill this void and began to study the literature.

As soon as I returned to the United States from Colombia, I went to the local library and bookstores to find books to increase my meager knowledge. To my surprise, there were very few. And in recalling my undergraduate and graduate studies in the earth sciences, I could not remember that any attention had been paid to the Moon or the Earth-Moon system. The first book I bought was The Measure of the Moon, by Ralph B. Baldwin.1 It turned out to be a fortuitous choice. Not only had Baldwin done a comprehensive survey of the literature (the specialized literature was much more extensive than that found in general bookstores), he had organized the existing knowledge and theories and pre­sented them in a readable fashion. His opening sentence was prophetic: ‘‘Every investigation of the Moon raises more problems than it solves.’’ During the next five or six years I would find myself immersed in these problems and dealing daily with the various protagonists cited in the research. I later learned that I was in good company by being impressed by Baldwin’s work; Harold Urey, a Nobel laureate in chemistry, had become fascinated by the Moon’s many myste­ries after reading Baldwin’s earlier book, The Face of the Moon, and had put forth his own theories on how the Moon formed.

My first impression that there was little source material quickly changed. Baldwin’s references were extensive, too many—in light of my new duties—for more than a cursory review. I settled on purchasing a few texts to read in their entirety and keep available as a small reference library. In addition to Baldwin I read The Moon, by Zdenak Kopal and Zdenka Mikhalov; Structure of the Moon’s Surface by Gilbert Fielder; Harold Urey’s The Planets and several of his articles

and reports; Gerard P. Kuiper’s ‘‘On the Origin of Lunar Surface Features’’; and an article by my old mines professor L. W. LeRoy, ‘‘Lunar Features and Lunar Problems.’’2

Perhaps most interesting of all, I discovered that most of the leading figures in lunar and planetary science, including Urey, Kuiper, Fielder, Kopal, and Baldwin, were active and accessible. In addition, some of the younger lions, such as Shoemaker, Frank Press at Lamont-Doherty, Jack Green at North American Aviation, John O’Keefe at the Goddard Space Flight Center, and Carl Sagan of Cornell University, were already involved in NASA programs.

The origin and age of the Moon had intrigued astronomers and Earth scientists for many centuries, with theories proposed based on a minimum of hard data. By the early 1960s existing theories had become more sophisticated, supported by ever increasing observational data and, soon, by returns from several of NASA’s unmanned programs. Three theories on the Moon’s origin held sway: (1) the Moon and the Earth had formed more or less simultaneously from the same primordial cloud of debris surrounding the Sun; (2) the Moon had been separated from the Earth either through tidal movements or by the impact of another body (some would split this into two theories); and (3) the Moon had formed elsewhere in the solar system, and in its orbit around the Sun it had been captured by the Earth’s gravitational field in an early close encoun­ter. Based on the information then available, each of these theories could be supported or argued against depending on one’s point of view and which data one considered most critical. The date when any of these events took place was also conjectural, but it was generally believed that the Moon had become Earth’s companion early in the formation of the solar system, some 4.5 to 5 billion years ago.

Certain information was well documented. The Moon’s physical dimensions and mass, its distance from Earth, and several other properties were known rather precisely. Unlike Earth’s, the Moon’s magnetic field, if any, was thought to be weak; its mass and volume translated to a body less dense than Earth, probably without an iron core or at best with a very small core. It had no discernible atmosphere. We knew that the Moon was locked into a slowly expanding orbit that allowed only one side to face Earth. The Moon’s farside or back side (not ‘‘dark side’’ as so many ill-informed writers call it, since it is lit by the Sun in the same manner as the side facing Earth) was a total mystery; was it the same as what we could see or very different? This lack of information had made the Moon’s farside the playground of science fiction writers for many years. One could imagine all sorts of strange things back there, including alien colonies.3

Probably the most contentious issue was the origin of lunar craters. Were they formed by some internal process like volcanism or by the impacts of small to large bodies like meteorites? The literature was full of this particular contro­versy, and the debate—at times vitriolic—went on at all lunar symposia. Each side had its champions, although it appeared that the “impactors” were begin­ning to win the day. Any of the three lunar origin theories could accommodate either an impact or a volcanic explanation, but the subsequent history or postorigin modification of the Moon’s surface would be entirely different de­pending on which crater theory proved correct. If the craters were volcanic, then the Moon’s interior had been molten after its formation and we could expect to find many Earthlike conditions. If the craters were caused by impacts, then the Moon’s evolution might have been very different from Earth’s, even though most students believed that impacts were common in Earth’s early history. Complicating this debate, we could observe other features on the Moon such as sinuous, riverlike rills and odd-shaped depressions that did not con­form to the contours expected of impact craters. What was the Moon trying to tell us? Had there once been water on the Moon? Had a combination of pro­cesses taken place? Were they still taking place?

A primary scientific justification for studying the Moon, with either manned or unmanned spacecraft, was to help us unravel Earth’s early history. A new term had been coined for such study, ‘‘comparative planetology,’’ and we used it frequently in our briefings both inside and outside NASA. Comparative plan­etology means studying the planets by comparing what can be observed or measured on one with similar characteristics on another; through this back – and-forth association we would increase our overall understanding of all the planets. We believed that applying this technique to the Earth-Moon system would be especially fruitful. In all the solar system, our Moon is the largest relative to the size of the planet it orbits—in essence we are a two-planet sys­tem. By studying the Moon we believed we would learn much about Earth. When the Apollo project began many basic questions concerning our home planet were unanswered, and many were similar to those we were asking about the Moon. How was it formed, and how had it changed during its early evolu­tion? What is the thick zone just beneath Earth’s crust—the mantle—made of?

How does the mantle influence or produce the energy that moves large sections of Earth’s surface?4

Earth’s surface is a dynamic place. Mountains rise and are eroded away, sea basins and lakes fill and dry up, and continents move vast distances, a process called “continental drift.’’ The record can be deciphered by earth scientists in the rocks of Earth’s crust. But our understanding becomes sketchier and more uncertain as we go back in time toward Earth’s earliest history. That part is obscured, hidden, or even destroyed by the very processes mentioned above. The oldest Earth rocks that have been positively dated, from northern Canada, are approximately 4 billion years old. The oldest piece of the solar system dated thus far is the Allende meteorite, calculated to be almost 4.6 billion years old, supporting the earlier theories that the solar system might be 5 billion years old. These dates, however, leave a gap of almost a billion years from the oldest dated Earth rocks to the solar system’s birth. This billion-year gap continues to be an enticing field for speculation and investigation.

Returning now to the three theories of lunar origin: What were their im­plications for Apollo? Could we expect to shed light on these riddles or perhaps even solve them? If either of the first two was correct—if the Earth and the Moon formed simultaneously and close together or if the Moon broke off from Earth, then one would expect the rock types or minerals we would find on the Moon to be similar to those on Earth. If the third theory was correct, that the Moon formed somewhere else in the solar system and was later captured by Earth, then we might find different rock types and minerals on the Moon, perhaps similar to some of the more exotic meteorites that have been recovered at various places on Earth. Regardless of the ultimate answer, we were confident we would be able to date the rocks and get a handle on a pressing question: When was the Moon formed? Some believed the Moon’s surface was ancient, that all the features we observed had formed early in its history and had changed little since then. Confirming this would be exciting; the Moon, as many were fond of saying, could act as a Rosetta Stone in deciphering the birth of the Earth and the solar system!

Harold Urey at the University of California, San Diego, was a strong propo­nent of the third theory. He believed the Moon had been formed through the accretion of planetesimals (large pieces of the primordial cloud from which the Sun and eventually the whole solar system evolved) and that this happened some 4.5 billion years ago. If true, it was an ancient and unchanged body and worthy of careful study. The Moon has an irregular shape (it is not a perfect spheroid but has slight polar flattening and an Earth-facing equatorial bulge), and it wobbles on its axis. Urey argued that the Moon had never been com­pletely molten or these irregularities would not have survived. According to his calculations, the Moon had formed as a somewhat cold body—those who said the maria were lava flows erupting from a molten interior were wrong. The maria, he believed, were the result of large-scale melting caused by the impact of large bodies, such as the one that had formed Mare Imbrium, and the maria material might have been the melted remains of carbonaceous chondrites, an unusual type of meteorite occasionally found on Earth. Urey was looking for­ward with great anticipation to obtaining lunar samples, especially from the maria (they should not be Earthlike lava), to prove his theory.

Urey’s reputation as a Nobel laureate was important in legitimizing our lunar studies. When he spoke, everyone listened. Although he had many differ­ences with other lunar scholars, sometimes he agreed with them. He agreed, for instance, that most craters were certainly of impact origin and that much of the lunar topography was shaped by ejecta from the impacts. He did not think there had been much volcanism on the Moon, but he accepted the observations of some volcanolike features. In a letter to Jay Holmes at NASA headquarters in January 1964 Urey said: ‘‘I am sure that only the most experienced hard rock geologist could possibly do anything about the subject satisfactorily. I urge strongly that all astronauts be well trained hard rock geologists. The Apollo project is being severely criticized by outstanding people, and I believe that if we do not at least [do] the very best that we can to solve important scientific problems that this criticism may well swell to a very great chorus.’’5 Urey’s suggestion on astronaut training was one of the first shots in a long campaign that led to the scientist-astronaut selections discussed in later chapters. Regard­less of his opinions, his presence at any lunar symposium guaranteed vigorous debate and lots of publicity, a commodity we eagerly sought as we struggled to make NASA management recognize how important the Moon would be in resolving issues of such magnitude.

Another vigorous debater was Thomas Gold, a professor at Cornell Univer­sity who had made his early reputation in astronomy. In recent years he had focused on problems related to the Moon. Tommy Gold was to prove a thorn in our sides with his strange theories, seldom supported by anyone else in the scientific community. His most controversial one, first proposed in 1955, was that the lunar surface was covered by a layer of fine particles eroded from the lunar highlands, perhaps several kilometers thick, that could move across the lunar surface and fill in depressions.6

He sought to prove this contention with photographs showing that most lunar features had a smooth appearance and many craters seemed to be filled rather uniformly with some material. He generally discounted the idea that this fill might have been molten material like lava or ejecta from impacts. Radar studies of the Moon tended to support his thesis that the uppermost soil layer was fine grained and of low density, but how thick this layer might be and what area it covered could not be resolved from the radar data.7 Other interpreta­tions were also possible.

The character of the lunar soil, especially its topmost layer, was of course a great concern, since it would directly affect the design of the lunar module (LM) and the astronauts’ ability to land and move around on the surface. Not much was known about how soils and fine-grained material would behave in the high vacuum found on the Moon. Several government and private labora­tories had done experiments to examine this question. Bruce Hapke at Cornell University, for example, had shown that fine particles deposited in a vacuum tended to stick together loosely, forming what he called ‘‘fairy castle’’ structures, or soils with low bearing strength.8 This could be seen as substantiating Gold’s contention of a low density lunar surface.

Before the return of pictures from Ranger, and later the Surveyor and Lunar Orbiter missions, photographs of the Moon had come from telescopic images, with a resolution of at best a thousand feet. Under such low resolution, every feature on the Moon appeared somewhat smooth. This problem did not deter Gold. Even after we received the higher resolution Ranger, Surveyor, and Lunar Orbiter photos, he continued to predict that when the lunar module attempted to land it would sink out of sight in his electrostatically levitated dust. At this early stage such predictions alarmed NASA’s engineers, for it was difficult to prove him wrong.

Fortunately questions of this type—though not so outrageous—had been anticipated, and the Surveyor spacecraft were designed to answer them. Sur­veyor did prove Gold wrong, which he accepted grudgingly, continuing to maintain that some areas of the Moon were covered with fluffy dust. He clearly enjoyed being the center of controversy, and after Surveyor’s deflator he came up with another whopper: the lunar dust would be pyrophoric. When the astronauts landed and opened their LM hatch, the oxygen released from the cabin would combine with the soil and cause an explosion. His reasoning was that the lunar surface, exposed for eons to the bombardment of the solar wind, had become oxygen deficient and would undergo an explosive oxidation when exposed to the LM atmosphere. This prediction also worried the engineers, and it would not be possible to prove or disprove it with any projects in the pipeline before the actual landing.

The school of volcanic crater supporters started strongly and slowly declined in influence as more and more observational and experimental data became available. But in 1963 and 1964 they still made a good case for their views. The leaders of this school were Gerard Kuiper, at that time director of the Lunar and Planetary Laboratory in Tucson, John O’Keefe at Goddard Space Flight Center (GSFC), and North American Aviation’s Jack Green. Each of these advocates had a somewhat different interpretation of what was observed on the Moon. Both Kuiper and O’Keefe admitted that impacts had played a role in the Moon’s evolution, but they still thought volcanism was the major explanation of its present surface formations. Kuiper had been an early student of the Moon. Ignoring Urey’s counterarguments, he believed the original substance that came together to form the Moon contained enough radioactive material to eventually raise the interior temperature and melt the entire Moon. In his model this had occurred some 4.5 billion years ago, forming the maria and filling the larger craters, all subsequently modified by meteoroids.

Green, however, took a hard-line approach. Essentially all features on the Moon could be, and should be, explained by volcanic processes. Jack was a colorful figure, never taken aback by criticism, and a good debater. You could count on him to enliven any lunar symposium. His forte was showing side-by­side photographs of terrestrial and lunar features that looked almost identical. The terrestrial features, of course, were always volcanic in origin.

The impact school was led by Gene Shoemaker and his United States Geo­logical Survey (USGS) followers. Gene had been influenced by an earlier and revered USGS chief geologist, Grove K. Gilbert, who in 1893 published a paper concluding that the Moon’s craters were probably of impact origin.9 Gene had carefully studied Meteor Crater in Arizona, just east of his new Flagstaff offices, as well as several other craters of known impact origin in other parts of the world. Robert Bryson, from NASA headquarters, had funded Gene to develop a detailed report of his findings that would combine his earlier studies and field observations at Meteor Crater. By 1964 Gene’s studies had been completed for some time, but he had not finished the written report. This was a sore point with Bob because so little had been published on the geology and mechanics of impact craters, and Gene’s work was intended to fill this void. He had published a short report on his work in 1963, but the full report was still in draft form.10

Bob, a former USGS geologist, had great insight into what it would take to convince the scientific community that important information could come from lunar studies. In addition to Gene’s work, Bob funded some of the studies of Ed Chao at USGS, who in 1960 discovered coesite in the shocked debris from Meteor Crater, a type of silica that forms only under extremely high pressure. Before Chao’s discovery, coesite had been made in the laboratory but had never been found in nature. This mineral is now a key fingerprint for identifying impact craters. Soon after this discovery Chao found stishovite, another form of high pressure silica, in rocks ejected from Meteor Crater-further confirma­tion that an impact of enormous energy had created the crater. Chao was later detailed to NASA as Apollo science work expanded, and we worked together under Will Foster. Bryson also funded the telescopic mapping of the Moon, initially through Robert Hackman at USGS. These maps laid the groundwork for all the subsequent lunar geological interpretations used during the Apollo landings and the planning that preceded them.

Despite the annoyance at NASA headquarters about the Meteor Crater re­port, Gene was a walking encyclopedia concerning what happens when a rela­tively small meteorite hits a solid object like Earth. (The iron meteorite named the Canyon Diablo that blasted the four-thousand-foot-diameter Meteor Cra­ter probably weighed about seven thousand tons.) He extrapolated these results to the larger lunar craters that must have been formed by even larger bodies. He was joined in this knowledge by experimenters such as Donald Gault at the NASA Ames Research Center and others who had conducted small hyper­velocity, laboratory-scale impact studies. In addition to making direct field observations on Earth, Gene and his staff, following Bob Hackman’s lead, had spent considerable time mapping the Moon using several large telescopes. Ap­plying standard terrestrial geological interpretations to these eyeball studies, they had become convinced that the Moon was pockmarked with impact cra­ters. Shoemaker was sure that almost all lunar craters had been formed by this mechanism, not through volcanism.

In a trip report of a visit to Menlo Park in May 1963, Bob Fudali described his conversations with Henry Moore, Dick Eggelton, Donald Wilhelms, Harold Masursky, and Michael Carr of USGS.11 After spending many hours drawing geological maps of the Moon based on telescopic observations, the USGS geolo­gists believed that, despite the high density of impact craters, there was substan­tial evidence of volcanic activity on the Moon, somewhat at variance with Shoemaker’s views. They also believed there was evidence that the maria were covered with extrusive igneous material, and they were convinced that tektites (rounded glassy bodies probably of meteoritic origin found at several places on Earth) originated on the Moon, thus supporting O’Keefe’s theories. Because of the chemical composition of the tektites, this meant that at least some parts of the Moon were ‘‘granitic,’’ which in turn meant that at some point in its evolu­tion the Moon had undergone differentiation in the presence of water. One could then conclude that the Moon was at least somewhat like Earth.

In addition to these major theories and vigorous debates, several related questions had puzzled lunar scientists for many years. Answers were especially important to the new breed of comparative planetologists, for they hoped the answers would shed new light on similar questions about Earth’s evolution.

During its early formation, Earth went through partial melting and differen­tiation. As the material that was to make up the bulk of Earth’s mass accumu­lated, the heavier material sank to the center, forming a core. Each layer above the core was of decreasing density, and the lightest materials formed the crust. Although we do not completely understand these various deep materials that form the bulk of Earth’s interior, we can infer and calculate what they are. Based on this knowledge, we have reconstructed the processes that formed them. As an example, we know that Earth’s continents are relatively light material ‘‘float­ing’’ on denser underlying rock. We also know that through geologic time there has been a constant churning of the upper layers and that Earth’s surface today looks very different than it did, say, 3 billion years ago. Although we say we know these things, they are really just theories based on observable field data and hypothetical calculations. It would be reassuring if we could find other examples of these processes or similar ones elsewhere in the solar system. What better place to look than the Moon, our closest neighbor?

Had the Moon undergone differentiation in its early history? Telescopes showed mountains on the Moon. They were generally lighter in color than the lowland maria and thus probably different in composition. Were the moun­tains less dense, as terrestrial mountains are less dense, on average, than Earth’s crust and upper mantle? If you believed that tektites came from the Moon, differentiation was a given, with less dense material occurring at the surface. Did the Moon have a core? The tiny but measurable magnetic field (averaging five gammas and believed to be due primarily to the interaction of the Moon with the solar wind) and overall lower density seemed to negate a lunarwide field, but we had not been able to make close-up measurements. Perhaps there were weak, relict local magnetic fields that would be evidence of early core formation. Why did the nearside and farside of the Moon look different? This question became more important when we received Lunar Orbiter pictures of the Moon’s farside with much higher resolution than those returned by Lunik 3 and the full extent of these differences became known. Did Earth-Moon tidal effects account for these differences, or was it some other factor?

Whether water ever existed on the Moon was another important question. Because the Moon has no discernible atmosphere (it was estimated to be equiv­alent to Earth’s atmosphere at altitudes above six hundred miles, appropriately an exosphere),12 water probably would not be found on the lunar surface under any conditions, but it might still exist belowground. Some proposed that it might be found in permanently shadowed craters near the lunar poles. Urey in 1952 and Kenneth Watson, Bruce Murray, and Harrison Brown in 1961 pro­vided an analytical basis for such predictions. The latter authors concluded, ‘‘In any event, local concentrations of ice on the moon would appear to be well within the realm of possibility. Unfortunately, if it exists, it will be found in shaded areas, and attempts to determine whether it is present must await the time when suitable instruments can be placed in those areas.’’13 Some thirty-five years later the Clementine and Lunar Prospector missions seem to support their analysis, though it is probably safe to say the authors had not imagined that ice would be detected by instruments in lunar orbit; such a possibility was beyond their dreams in the early 1950s.

On Earth, water is needed to form granites, so if granites existed on the Moon, then water must have been present in its early history. If water could be found on the Moon it would greatly simplify our plans for post-Apollo manned exploration. Its presence in an easily recoverable form would reduce the potable water we would have to transport to the Moon, and water could be used as a source of oxygen for manned habitats. Far-out planners even envisioned mak­ing rocket fuel by separating the hydrogen and oxygen. The questions posed by present-day space planners or raised by the information gained from the Clem­entine and Lunar Prospector missions thus are not new but were on our minds thirty years earlier.

Would we find any evidence of life forms, however primitive, in the samples brought back to Earth? This outcome was considered unlikely but not impossi­ble. For this reason the samples and the astronauts would be quarantined on their return lest they carry some deadly virus or pathogen to which we poor earthlings would have no immunity. Any evidence of life would be astounding and would require rethinking how life formed on Earth.

All the questions above, and their answers, were important both to NASA (especially my office) and to the scientific community in general. Our post – Apollo mission strategies were based on attempting to find answers, which in turn would help us plan our programs for Venus, Mars, and beyond, using the Moon as a staging point for these more difficult missions. And there was still the link to understanding Earth.14

All these theories, questions, and debates could be resolved by a relatively small suite of activities and experiments. The trick would be to design them so they could be carried on the missions and deployed by the astronauts. The astronauts would have to sample the rocks and soil at their landing sites over as large an area as possible and bring the samples back to Earth for analysis and reconstruction of their geological context. Also, to complete the picture they would need to carry certain geophysical instruments to collect data pertaining to the Moon’s subsurface or other environmental conditions. In the introduc­tion to his book, Baldwin had stated: ‘‘It is beyond hope that we shall ever have a complete and definitive answer to all lunar problems.’’ Finally he had predicted: ‘‘Landing on the moon and analyzing its materials will help greatly but will raise more problems than are solved.’’15 These predictions echoed concerns raised in his first chapter. We hoped that our plans for extensive manned lunar explora­tion would go a long way toward changing his mind on both of them.

After becoming reasonably familiar with the current state of knowledge about the Moon, I started making some personal observations. I got permission from Tom Evans to contract with the Astronomy Department at the University of Virginia for time on their large (twenty-six-inch refractor) telescope so some of us on the NASA headquarters staff could travel to Charlottesville and make our own close-up studies. Laurence Fredrick, director of the Leander McCor­mick Observatory, was a gracious host for those of us that took advantage of the opportunity. This telescope, almost a twin to the famous Naval Observatory telescope in Washington, D. C., where some of the first lunar studies had taken place in the nineteenth century, including those by Gilbert, was the one USGS used in 1961 to begin the detailed mapping of the Moon funded by Bob Bryson. Because this work had recently been transferred to the Lick Observatory in California and a new observatory near Flagstaff, observing time was available. The Virginia telescope was an ideal instrument for casual Moon viewing be­cause with easily mastered techniques it provided a resolution of a few thou­sand feet for lunar surface features. Charlottesville was only a two-hour drive from Washington, so we could leave the office immediately after work, stop for a quick dinner, set up the telescope in plenty of time for a few hours of viewing, and still get home shortly after midnight.

A twenty-six-inch-refractor telescope is a very large piece of equipment. The telescope with its mount weighed some eight tons. A rotating dome with sliding doors covered the telescope, and housed within the dome were the electronics and motors that allowed one to point and track the telescope. Under Larry Fredrick’s tutelage, I became adept at operating the instrument, and after a few nights’ practice I was able to observe by myself. As one might expect, viewing was ideal on clear nights, and the winter months were best of all because cold, stable air reduces atmospheric disturbances. But even on exceptionally clear nights there was always a shimmering distortion caused by Earth’s atmosphere, making it appear that heat waves were rising from the Moon and tending to obscure features under high magnification. I spent many a cold night studying the Moon’s surface, following the terminator as it slowly moved across the face of the Moon revealing the surface detail. When the Sun’s angle was correct I could compare my observations with the first USGS lunar maps of the Coper­nicus and Kepler regions to understand how this latest attempt to map the Moon geologically was carried out and why the USGS mappers were identifying certain types of surface features as discrete geological formations. The subtlety of most of these features was evident, and I came to appreciate how an earth – bound geologist’s imagination might become a dominant factor in drawing a geological map of the Moon with the enormous disadvantage of never having set foot on the surface.

Another compelling reason for spending time observing the Moon was the recent spate of reports by reputable astronomers about transient phenomena on the lunar surface. In 1958 a sensational announcement had been made by Soviet astronomer Nikolai Kozyrev, who claimed he had recorded spectra of a transient event on the Moon near the central peak of the crater Alphonsus. Other observers soon reported color changes and similar events at other lunar features, the most exciting being at the crater Aristarchus.

Excerpts from the report written by James Greenacre, employed at that time by the U. S. Air Force Lunar Mapping Program at Lowell Observatory near Flagstaff, Arizona, tell his exciting story of what he observed one night at Aristarchus.

Early in the evening of October 29, 1963, Mr. Edward Barr and I had started our regular lunar observations. . . . When I started to observe at 1830 MST. . .

I concentrated on the Cobra Head of Schroeter’s Valley. . . . at 1850 MST I noticed a reddish-orange color over the dome-like structure on the southwest side of the Cobra Head. Almost simultaneously I observed a small spot of the same color on a hilltop across the valley. Within two minutes these colors had become quite brilliant and had considerable sparkle. I immediately called Mr. Barr to share this observation with me. His first impression of the color was a dark orange. No other color spots were noted until 1855 MST when I ob­served an elongated streaked pink color along the southwest rim of Aristar­chus. . . . at approximately 1900 MST I noticed the spots of color at the Cobra Head and on the hill across the valley had changed to a light ruby red. . . . I had the impression that I was looking into a large polished gem ruby but could not see through it. Mr. Barr’s impression of the color at this time was that it was a little more dense than I had described it. . . . By 1905 MST it was apparent that the color was fading.16

Greenacre and Barr did not advance any theories on what may have caused the colors they observed, but in a contemporaneous report John Hall, director of the Lowell Observatory, vouched for the authenticity of the sighting. He called Greenacre ‘‘a very cautious observer’’ and noted that Greenacre’s boss, William Cannell, ‘‘stated that he could not recall that Greenacre had ever plotted a lunar feature which was not later confirmed by another observer.’’17

Thus was reported the first sighting of a lunar transient event, confirmed by two observers and, most important, made by highly qualified personnel. A second sighting by Barr and Greenacre, at the same location, was recorded one lunar month later on November 27, 1963.18 This observation also was con­firmed by Hall and by Fred Dungan, a scientific illustrator on the staff and a qualified telescopic observer. This color feature was reported to be somewhat larger than the one observed in October. It seemed beyond a doubt that some­thing was going on near Aristarchus, since other observers before and after Greenacre and Barr recorded similar activity in the vicinity.

Aristarchus is the brightest feature on the Moon’s nearside. This fact, along with the odd shapes of nearby features, suggested that it was of ‘‘recent’’ vol­canic origin. (Recent is a subjective term, since no one could then be sure of the relative ages of any lunar features, and the absolute times when they were formed were even larger unknowns.) By USGS’s reckoning brightness equated to ‘‘young,’’ and these color changes could mean that volcanic processes were still taking place on the Moon. This was an exciting prospect for those of us deciding what experiments to perform on the Moon. Thus, every night that I spent at the telescope I devoted some time to looking at Aristarchus, hoping I would see one of these ‘‘eruptions.’’ I never did.

After setting up the contract at the University of Virginia, I contacted an astronomer friend at the NASA Goddard Space Flight Center, Winifred C. ‘‘Wini’’ Cameron, suggesting we start a nationwide network of amateur and professional astronomers to maintain a continuous Moon watch for transient phenomena. Wini was already studying the origin of lunar features and was working with John O’Keefe at GSFC, so this activity fit neatly with her ongoing work. The idea was to publicize a telephone number where people could call in their observations. The person manning the hot line would then contact other observers to try to confirm the sighting. In spite of the acknowledged profes­sionalism of some who had made sightings, many in the small lunar commu­nity were skeptical about such events, so we needed to get independent confir­mation. We activated the network under Wini’s direction in 1965. She went on to study, extensively, lunar transient phenomena and began a program called Moon Blink that developed instrumentation specifically designed to measure and record such transient events.

Lunar transient events had been reported long before the start of the Apollo program, but as might be expected, Apollo aroused great interest in the Moon in amateur and professional astronomers alike. Many more reports of various types of sightings such as color changes, obscurations, and sudden bright spots were made after Apollo Moon landings became the centerpiece of NASA’s space program.19 Up until this time, however, except for Greenacre’s sighting, confir­mation had never been possible; subsequently there was independent confirma­tion of several events.

In 1967, after careful analysis of Lunar Orbiter У high resolution photo­graphs of the region of Aristarchus, scientists at the Lunar and Planetary Labo­ratory at the University of Arizona discovered some interesting features at the location of Greenacre’s color sightings. They reported that in Schroeters Valley, near the crater named the Cobra Head, they observed a volcanic-looking cone with flow features on its flanks, and that the crater Aristarchus showed evidence of volcanic activity.20 These discoveries suggested that Greenacre was observing the effects of ongoing lunar eruptions.

The information gained later during Project Apollo and from follow-on studies makes it seem likely that some type of gaseous emission or other surface changes did take place during this time. Some of the color changes reported may have been imagined or caused by terrestrial atmospheric distortion that fooled the observers, but some were almost certainly real events. Astronauts’ observations pertaining to lunar transient phenomena are discussed further in chapter 13. For more on the subject, see selected works by Cameron.21

What Do We Do after Apollo?

Even before we made detailed plans for including science on the Apollo mis­sions, we undertook planning and analysis for missions that would come later. When I joined NASA in 1963, this planning was being done in Tom Evans’s office under the name Apollo Logistics Support System (ALSS), implying a program that would come after the Apollo missions but would capitalize on the Apollo hardware then being designed. Post-Apollo programs were given other names in later years as management attempted to get a commitment to con­tinue lunar missions after the initial Apollo landings.

By late 1963, except for the effort that went into the Sonett Report, little had been done to fill the void in Apollo science planning. And many in NASA claimed that no void existed. The Apollo program had only one objective: to land men on the Moon and return them safely. The astronauts would probably take a few pictures, though no camera had yet been selected. They might pick up a few rocks, but tools for doing this were not under development, nor were we designing the special boxes essential for storing such samples on the return trip. A few forward-looking scientists were beginning to think about these con­cerns, but no one was receiving NASA funds to develop the equipment needed. Post-Apollo planning was an entirely different matter. Tom Evans’s office was already spending NASA funds to address what we should do on the Moon after the initial landings. His group and others in Advanced Manned Missions who were looking ahead had initiated studies at the Marshall Space Flight Center (MSFC) that led to the ten-volume MSFC report Lunar Logistic System. This effort was directed at MSFC by Joseph de Fries of the Aero Astrodynamics Laboratory, but it included contributions from other MSFC organizations.

In the fall of 1963, less than six years before the first Apollo Moon landing would take place, no timelines had yet been developed to tell us how long the astronauts would, or could, stay on the lunar surface. Payload numbers for the science equipment were not firmed up and varied from the 100 to 200 pounds estimated for the Sonett Report to the ‘‘back of the envelope’’ 250 pounds allotted later. We all assumed it would be difficult to get a larger allocation until all the Apollo systems had been tested and flown and had their performance evaluated. In spite of the many uncertainties and the lack of firm numbers, we took it as given that the landings (number undefined) would be successful and that the myriad Apollo systems would function as advertised.

Our job was not to question any of the Apollo assumptions. Another office in Advanced Manned Missions, under the rubric of supporting research and technology, was responsible for developing alternative ways to ensure mission success. Not only did we assume success, we were charged with expanding the capabilities of the basic Apollo hardware far beyond the original intent. For example, how could we upgrade the lunar excursion module (LEM) to carry a much larger payload than currently planned? How could we extend the time that the command and service module (CSM) could stay in lunar orbit? How could we increase the potential landing area accessible to the LEM (restricted for the first landings to the Moon’s nearside, central longitude, equatorial re­gion) so that we could explore what appeared to be critical geological sites far from the planned Apollo landing zone? And would it be possible to land a modified, automated LEM, turning it into a cargo carrier (LEM truck) in order to bring large scientific and logistics payloads to the Moon? All these questions and many more were already under study when I joined the office. (Later in the program the term lunar excursion module was shortened to lunar module, LM, but at this time LEM was still the preferred name.)

The missing ingredient in all this planning was an explanation for why we wanted to stay longer on the lunar surface and why we needed to modify the Apollo hardware to carry bigger payloads. How long should we stay? How big a payload? It became my job to get answers from the ongoing studies. At the end of July 1963, as one of his last actions at headquarters, Gene Shoemaker had sent a letter to Wernher von Braun, the Marshall Space Flight Center director, asking MSFC to suggest what types of scientific activities should be undertaken on the ALSS missions. Verne Fryklund, as Shoemaker’s successor at NASA, continued this effort, and I in turn inherited this inquiry when I informally joined his staff.

After meeting Paul Lowman in Fryklund’s office, I quickly learned that he shared my enthusiasm about studying and exploring the Moon. Not having been exposed to normal Washington turf battles and jealousies, it seemed quite natural that I ask Paul to work with me informally on some of the projects I had begun. Paul had already made a name for himself by convincing the Mercury astronauts to use Hasselblad cameras on their flights to photograph the Earth’s surface. This was no mean accomplishment, since these former test pilots were much more interested in flying and monitoring spacecraft systems than in being photographers. Most of the astronauts eventually enjoyed taking photos, especially when they were published extensively in newspapers and magazines. At that time Life had an exclusive agreement with the astronauts to publish first-person accounts of the missions, and a few beautiful full-color photos of the Earth appeared in the articles that followed each Mercury flight. As a result of this success, Paul continued to coach the upcoming Gemini astronauts in photography.

One of the attractive aspects of working at NASA in those early days was that staff members were given great freedom to attack whatever problem they un­covered, without bureaucratic red tape and worry about turf. Paul had orig­inally accepted his temporary headquarters assignment in order to work with Gene Shoemaker, so with Gene’s departure, the reorganization of Fryklund’s office, and the arrival of Will Foster, the timing was right. Thus we began a long professional friendship that endures today.

By the time I joined Evans’s small team in 1963, we already had the results of some preliminary studies on expanding the versatility of the Apollo hardware. The MSFC Lunar Logistic System study had examined the hardware then under development for Apollo and documented its inherent flexibility. With what we claimed would be minor modifications, it would be possible to land the LEM at selected sites with no crew on board. Such a LEM could then be a cargo ship carrying as much as seven thousand pounds to the lunar surface, replacing ascent fuel and other equipment not needed for a one-way, unmanned trip. A LEM with this capacity could carry living quarters, large science payloads, or other types of equipment depending on the mission. It seemed that a crew of two astronauts, arriving in another modified LEM and landing close to one or more unmanned logistics LEMs, could spend as much as two weeks on the Moon by either transferring to the earlier-landed LEM or using other payloads that had preceded them.

Similar studies of the CSM showed that it could be kept in lunar orbit long enough to support a two-week lunar stay. In addition, remote-sensing payloads could be carried in one of the CSM’s bays to map the lunar surface in various parts of the electromagnetic spectrum, an undertaking that was receiving more and more backing and attention.

Most of my office colleagues were engineers with degrees in electrical, aero­nautical, or mechanical engineering and little training in earth sciences. This background was mirrored by NASA’s senior management. We decided the best way to convince our bosses that there would be exciting and important inves­tigations for the astronauts to undertake on the Moon (requiring many days and a wide variety of equipment) would be to illustrate these tasks with ter­restrial analogies and describe the type of fieldwork and experiments required on Earth to unravel its own history.

Drawing on the Sonett Report and our own knowledge and experience, Paul and I first visited the rock collection at the Smithsonian Museum of Natural History. We borrowed rock samples of various types that illustrated the Earth’s geological diversity and the complex geological and geophysical situations we believed would be encountered on the Moon. With visible evidence of how a planetary body (the Earth) had evolved, we developed a rudimentary ‘‘show and tell’’—a short course in terrestrial geology and geophysics for NASA deci­sion makers—and extrapolated this lesson to the Moon. We hoped our rock collection, along with maps, photos, cross sections, and such, would stimulate their interest and demonstrate that what we were proposing was real and im­portant. We selected igneous, metamorphic, and sedimentary rock samples, later augmented by a few specimens collected at Meteor Crater, Arizona, that showed how a meteorite impact could make rocks look much different than before they were struck. In 1963 so little was known of the physical characteris­tics of the lunar surface that we felt free to use almost any type of rock to tell our story. Armed with our teaching materials, we put together a half-hour lecture designed around passing out our rock collection to the audience to make particular points and—we hoped—elicit questions. We started with my office colleagues, honed the presentation, and later lectured to senior staff. Tom Evans and E. Z. Gray were impressed with the story we put together. We were ready to take our show on the road and present it along with recent study results con­firming that the astronauts might be able to stay on the Moon for two weeks deploying sophisticated science payloads.

On December 23, 1963, after just four months of getting our story together, Evans was asked to brief a prestigious audience: Nicholas E. Golovin, a member of the President’s Science Advisory Committee (PSAC), and staff from the Office of Science and Technology (OST). Golovin had been a senior manager at NASA before going to PSAC. He had earned a reputation as a stern, no­nonsense leader in NASA’s early days when he chaired a committee to review the Apollo launch vehicle options and became involved in the internal debate on selecting lunar orbit rendezvous (LOR) as the preferred mission mode. Tom was apprehensive about the briefing, which was designed to inform PSAC about our thinking on post-Apollo missions. Ed Andrews and I went with Tom, but because of Golovin’s reputation we were told just to listen unless Tom asked us to answer a question.

I thought the briefing went well, and I only responded to a few “geological” questions directed my way. Golovin asked several questions, some in a peremp­tory tone that I assumed was his normal manner. Donald Steininger, from OST, asked a few questions on classifying rocks, obviously trying to understand how much sampling would be necessary to understand the Moon’s history. Tom saw the meeting more negatively. He didn’t think we had convinced our audience of the need for extended lunar exploration. As it turned out, Tom’s instincts were right: after President Kennedy’s death, the Johnson administration never fully embraced post-Apollo lunar exploration.

Of course, not knowing in 1963 and 1964 what events would take place that might dash our plans, we charged ahead and prepared for the big show, a briefing on our vision of post-Apollo lunar exploration for George Mueller, Tom and E. Z. Gray’s boss. Mueller, a former professor of electrical engineering, was a slender man with dark hair combed straight back, whose thick, black- rimmed glasses gave him an owlish look. In the meetings I had attended he was soft-spoken and deliberative. I was looking forward to this chance to brief him. Mueller’s management style was somewhat unusual compared with that of other managers I had known, and in the years ahead it set the tone for the Apollo program.

After we moved to 600 Independence Avenue (across the street from a parking lot that later was the site of the Smithsonian Air and Space Museum), briefings and status reviews for Mueller were held in Office of Manned Space Flight (OMSF) conference room 425. The room was set up to hold forty to fifty, with Mueller and senior OMSF management seated in the front row before three back-projected screens. A lectern for the presenter was usually placed to the audience’s left of the screens. Several overhead microphones let the pre­senter prompt the projectionist for the next vugraph or slide. Al Zito, a civil servant transferred from the navy, ruled the seas behind the screens. You soon learned that if you wanted a smooth presentation, Al had to understand your needs. With an assistant, he would work the three screens like an orchestra conductor, never missing a beat even if the presenter lost his place or questions disrupted the flow. Al became an OMSF institution. He could have written a funny book about NASA in the years leading up to the first Apollo flights, for he was privy to more senior-level decision making than almost anyone else. Such a book could have included the faults, foibles, and stumbles of many senior managers unprepared for the grilling they got on the stage in room 425.

We had a small art department to develop presentation material for OMSF offices. Housed in the basement of 600 Independence Avenue, it was run by Peter Robinson, who had a full-time staff of six or seven artists and technicians. Pete was a true NASA treasure-unflappable in the face of impossible deadlines yet smiling and friendly and somehow always delivering the goods. I came to know Pete and his team well over the years. I often spent hours in Pete’s office along with Jay Holmes, who worked on Mueller’s staff to develop presentations, sketching and revising new material for briefing senior management. Mueller had a special ability to make a flawless presentation with minimum preparation before audiences of all descriptions, keeping them spellbound with the colorful and exciting pictures we and others provided. Every program manager soon learned to keep a file drawer full of up-to-date vugraphs of his project, ready at a moment’s notice to either give a presentation or provide material for someone else to present.

Although the conference room had microphones to cue the projectionist, there was no way to amplify what was being said for those in back. During and after presentations, Mueller and his staff would ask questions and discuss the matter at hand, with Mueller taking the lead. His voice was soft and low, and since he seldom raised it, even during contentious debates, everyone would be absolutely silent so as not to miss what was being said in the front of the room. In spite of straining to hear, those of us in the cheap seats often could not get the gist of the discussion.

After the meeting we would discreetly mill around in the corridor outside asking ‘‘What did he say?’’ about a particular subject of interest. We usually had to ask two or three people before we got the whole answer, since even those seated closer might not have heard everything. I have often wondered if Mueller knew about these sessions and purposely pitched his voice low to keep everyone focused and eliminate unwanted questions on his time. Whether or not it was a ploy, his meetings usually zipped along, unlike those run by many other man­agers I have worked with.

The staff had two strategies for briefing Mueller. During the regular work­week we tried to schedule our briefings early in the morning, because as the day wore on, even if you were on his schedule, he would often be called away for urgent telephone calls or for short or long discussions back in his office. His calendar was always filled, so if you didn’t finish your briefing in the time allotted it was difficult to get back on his agenda. We quickly learned to schedule important decision-making meetings on Saturday or Sunday, when interrup­tions were at a minimum and we could talk in a more relaxed environment. NASA Manned Space Flight under Mueller became a seven-day-a-week job, and the lights burned late in most offices at headquarters as we tried to keep up with the rapidly evolving program. The same was true, I know, at the NASA centers.

Our briefing for Mueller was carried out in an atmosphere less formal than usual and with fewer attendees. We made our case for longer staytimes and larger payloads, and since I was at the front for my presentation, this time I had no trouble hearing his questions. Our briefing and props succeeded beyond our expectations; eventually E. Z. Gray felt comfortable enough with our story that he borrowed our presentation for his own briefings, and Mueller soon began to lobby for post-Apollo missions. Over the next two years, as more and more in­formation on the Moon’s characteristics became available through new studies and the unmanned missions, we improved our story and eventually made our presentation, without the rocks, at national scientific meetings and symposia.

In the spring of 1964, as we continued to spread the gospel of lunar explora­tion, Tom Evans scheduled a trip to Houston to discuss our ideas and plans for post-Apollo exploration with some of the staff at the newly formed Manned Spacecraft Center (MSC; later named the Lyndon B. Johnson Space Center). Many of the new arrivals at MSC had been transferred from the NASA Langley Research Center, and one of the more senior was Maxime ‘‘Max’’ A. Faget. Max was a feisty aeronautical engineer who had been a member of the NASA Space Task Group, the source of many of the initial Project Mercury program man­agers and other senior managers for the fledgling NASA. In 1959 he served on the Goett Committee that recommended increasingly difficult missions, from Project Mercury to Mars-Venus landings, including manned lunar landings. With this background we thought he would be interested in and supportive of our plans. Max’s title was director of engineering and development, and as one of the designers of the Mercury capsule he now led the MSC engineering teams responsible for the design of everything from the LEM to space suits.

Tom took three of us with him to Houston to be available for questions from Max and whoever else he might invite to the briefing. At this time the MSC staff was still small. Some members, including Max, were housed in a building near downtown Houston while their permanent offices were being built in a cow pasture at Clear Lake, about twenty miles southeast of Houston. Max brought about six staff members to our briefing, which Tom Evans gave in its entirety. He described in detail the type of tasks we thought would be needed after the initial Apollo landings to answer fundamental questions about the Moon’s origin and explained the value of using the Moon as a lunar science base. To carry them out, Tom explained, would require making changes to the projected Apollo hardware so that astronauts could remain on the Moon for weeks at a time and so that large logistical payloads could be carried. As the briefing progressed, there were no questions from Max or any of his staff. Finally, after about an hour of talking, Tom completed the briefing and asked for comments or questions. After a short pause, Max, a short, stocky man with a receding hairline and a bulldog demeanor, turned in his swivel chair and asked in a raspy voice, of no one in particular, ‘‘Who thought up these ideas, some high-school student?’’

Despite his look of great consternation, Tom calmly tried to explain how we had arrived at our position, but it was clear that Max wasn’t interested. Perhaps he had more pressing matters on his mind, such as the first Gemini program launch, which would soon be announced. Perhaps he knew that these ideas were based in part on work done at MSFC, a rival for management of pieces of the Apollo program. The briefing ended in some disarray because of Max’s attitude. We quickly left and flew back to Washington, dismayed at our inability to get a more positive response. This was my first encounter with Max Faget and some of the MSC science staff, and it signaled the beginning of a long and often contentious relationship with some MSC offices that lasted until the final Apollo flight splashed down.

No story about NASA would be complete without some discussion of bud­gets. There have been several accounts, perhaps apocryphal, of how NASA administrator James Webb and his staff arrived at a dollar figure for how much the Apollo program would cost American taxpayers. The most common story had it that his managers told him it would take $12 billion or $13 billion to achieve a manned lunar landing and return, so he made an appointment to discuss the program and budget that he was recommending with President Kennedy. On the way to the White House in his Checkers limousine, a modified version of the popular taxicab (he was the only agency head to use such inele­gant transportation, which he found spacious and easy to get in and out of), based on his experience as director of the Bureau of the Budget and his exper­tise in dealing with big government programs, he doubled the estimate to $25 billion. Whether or not the genesis of this number is true, his projection was on the mark, and the Apollo program eventually was completed for almost pre­cisely that amount.

Webb and his deputy, Hugh Dryden, were the only political appointees at NASA. Webb had been appointed by President Kennedy at the beginning of his term to succeed NASA’s first administrator, T. Keith Glennan. Webb was a lawyer who came to NASA from the private sector, but he had been a senior government official in previous administrations and still maintained close ties to important political figures. During his tenure at NASA he was admired for his political astuteness and his ability to move Congress and administrations in the directions he chose. As the Mr. Outside of NASA, he smoothed the way for the agency to grow and prosper during the hectic first years of the Apollo era.

I don’t recall any meetings with Webb or Dryden—I was much too junior for such exalted company—but I did attend many meetings over the years with Bob Seamans, the associate administrator and number three man in the manage­ment pecking order. His background was very different from Webb’s. He had spent most of his career at MIT, first as a professor and later working on a variety of military projects at what was then called the Instrumentation Labora­tory. In his autobiography, Aiming at Targets,1 Seamans recounts being re­cruited by Glennan in 1960 to be NASA’s ‘‘general manager,’’ running the day – to-day operations. After Webb succeeded Glennan, Seamans continued to fill the general manager’s position and became NASA’s Mr. Inside. It was in that role that I first met him soon after I joined NASA. I’m sure he wouldn’t remem­ber that meeting, and I don’t recall the subject (although it probably had something to do with lunar exploration), but I remember one exchange vividly. During the presentations, I asked a few questions. Seamans turned abruptly in my direction and said in a pained voice, ‘‘This is my meeting.’’ I may not remember what was covered at the meeting, but those words are etched in my memory. His outburst quickly put a lowly GS-13 in his place, and from that point on I only listened.

Under Seamans’s direction NASA quickly became a polished management team. He instituted comprehensive monthly status reviews (general manage­ment status reviews) where he presided. Every aspect of all the programs was reviewed, problems were thrashed out, and actions were assigned. It was almost impossible to hide a problem in such a forum, and the business of the agency moved ahead briskly. Eventually Seamans was appointed deputy administrator, and he stayed at NASA until January 1968, the eve of Apollo’s biggest successes, for which he could take major credit. In 1974 President Gerald Ford appointed Seamans to lead a new government entity, the Energy Research and Develop­ment Agency, and I had the pleasure of working for him again, only this time in a much more senior role.

Only a small fraction of the $25 billion Webb asked for found its way into the Advanced Manned Missions budget or its predecessor offices. It has been diffi­cult, thirty-five years after the fact, to reconstruct these budgets from existing NASA documentation and from my own files. But it appears that from fiscal year 1961 to FY 1968 our offices received about $100 million out of the overall Manned Space Flight budget. These dollars funded a variety of studies: manned lunar and planetary missions, vehicle studies, Earth orbital missions, systems engineering, and other special studies, all related to programs that might follow a successful Apollo landing. In turn, Evans was allocated his small portion of these overall budgets for his office’s studies. By FY 1964 he had received a little over $7 million, which he had divided among five competing study areas, and increased funding came our way over the next few years. In the first two and a half years that I worked for Tom and his successors (calendar year 1963 to CY 1965), we had access to about $8 million to start obtaining some hard numbers that would back up the ‘‘how long, how big’’ assumptions for the ALSS missions that we grandly threw around in our briefings and rock lectures. In addition to contractor studies, this funding included a few hundred thousand dollars that was transferred to the United States Geographical Survey (USGS) in FY 1964 and FY 1965, to begin geological and geophysical field studies of how to carry out specific operations during lunar missions with long staytimes. In the early 1960s, you could get a lot of bang for your NASA buck.

My first contractor study was undertaken toward the end of 1963 by Martin Marietta. The company had been in competition with Grumman to build the lunar excursion module, and in the final selection Grumman won. During the competition, Martin had built a full-scale mock-up of its concept of what a LEM would look like. Not surprisingly, since they were both bidding to the same specifications, the Martin concept looked very similar to the winning Grumman model. This mock-up now sat in a high-bay building at the Martin plant in Middle River, Maryland, near Baltimore. Disappointed by the loss, and learning of our activities, a Martin manager came to my office one day to see if there was any interest in using this equipment. Having just completed a param­etric analysis of contingency experiments for Apollo, I saw the opportunity to determine, in a preliminary fashion, what difficulties the astronauts might have in making observations from the LEM once they landed on the lunar surface and before they set foot outside. In the back of our minds was the fear that after a successful touchdown something might keep them from getting out on the lunar surface.

Because Martin had the only look-alike version of a LEM, I was able to justify a sole-source contract, and one was soon in place. As part of the contract, Martin did its best, within our funding limitations, to simulate a lunar surface surrounding the LEM mock-up on the floor of the high-bay building. Tons of ashes, sand, and other material were poured on the floor, and we also scattered various types of rocks in the loose, finer-grained material, including some of those we had borrowed from the Smithsonian. To simulate lighting conditions the astronauts might encounter on the Moon, we illuminated the simulated surface with light ranging from low to intense and varied the angle to duplicate the changing sun angles they might confront depending on when during a lunar day they landed.

Since this was to be a simulation of human factors as much as geological conditions, the contract was managed by the Martin human factors department under the direction of Milton Grodsky. The “astronauts” were Martin em­ployees selected by the company. Paul Lowman and I gave them some rudimen­tary geological training, concentrating on how to make visual observations, provide verbal descriptions using geological terms, and take photographs from the LEM windows to show the nature of the simulated lunar surface. The

Martin test subjects volunteered to spend three or four days isolated in the LEM mock-up, eating and sleeping in the confined space and able to communicate with the test engineers only by radio. The living conditions inside the Martin mock-up, though somewhat uncomfortable, were considerably better than those faced by Neil A. Armstrong and Edwin E. ‘‘Buzz’’ Aldrin Jr. five years later during the first lunar landing and by astronauts in later missions. Armstrong and Aldrin, for example, didn’t get much rest during their twenty-hour stay. When they tried to sleep after returning to the LEM from extravehicular ac­tivity (EVA) on the surface, Armstrong had to rest on top of the motor casing of the ascent stage rocket, while Aldrin curled up in a confined space on the LEM’s floor. Neither slept soundly, and Armstrong perhaps not at all. We were easier on our test subjects; we gutted the interior of the mock-up, and each test ‘‘astronaut’’ had enough space to sleep on a thin mattress on the floor.

The first problem was how to photograph and describe the scene outside the LEM, which had only two small windows, both facing in about the same direction. With this limited view, less than half the lunar surface would be visible if the astronauts could not get out. The LEM also had an overhead hatch to allow them to enter it from the CSM while in lunar orbit, and in that hatch was a small window designed to permit star field sightings, if needed, to up­date the LEM’s guidance and navigation system. But on the lunar surface this window would face only the dark sky above the Moon. The LEM would be equipped with a small telescope that could be operated from inside to assist in the star sightings. We simulated opening the hatch on the lunar surface, with one of the test subjects standing in the opening to make observations. That worked quite well, and we were confident that if this was allowed we could get a good description of the landing site supplemented by panoramic photographs. But what if the astronauts couldn’t open the hatch or weren’t permitted to do so?

Perhaps we could adapt the telescope—design it to operate more like a periscope so they could scan the surface in all directions. Paul and I traveled to Boston to ask these questions at MIT’s Instrumentation Laboratory. The lab had the NASA contract to design the guidance and navigation control system for the CSM and LEM. The telescope was an integral part of the system, along with a sextant in the CSM. We spent the afternoon describing our Martin study and explaining the added value of designing the telescope so it could not only take star sightings but scan the surface and accept a handheld camera to let the astronauts photograph the full surface area of the landing site from within the LEM. The engineers thought this would be possible, but it would entail a major design change to the telescope. Since they were already having some trouble meeting contract objectives, we knew that asking for such a change, based on a perhaps unlikely contingency, went beyond our pay grade. I wrote a short report of our visit and then drafted a memo to George Mueller, for Homer Newell’s signature, requesting that modifications to the LEM periscope be con­sidered to permit terrain photography and visual observations of the lunar surface.2 I have no record of how this request was processed in OMSF, but the modifications were considered too extensive and costly, and the matter was dropped. We resurrected this idea some time later, but again it was not imple­mented, and fortunately such an instrument was never needed on any of the Apollo landing missions.

With the Martin Marietta contract under way, I started to lay plans for several other studies. The Sonett Report made it clear that we would need a geophysical station of undetermined design that could support five or six ex­periments. A drill that could extract core samples from deep below the lunar surface was another piece of equipment we believed the scientific community would eventually call for. After studying the first USGS geologic maps of the Kepler and Copernicus regions, traverses of tens of miles seemed necessary if we were to fully understand such large craters, some twenty and fifty miles in diameter. To work far beyond their immediate landing site, the astronauts would have to be mobile, and the more capable we could make a vehicle the more useful it would be. According to our limited understanding of the ongo­ing designs for the astronauts’ space suits and life-support backpacks, they would never be permitted to make such long traverses on foot; they would need a vehicle with a pressurized cab and full life support.

Our growing knowledge of the Moon suggested that the lunar surface might be stable, not subject to shaking and movement. If that was true, it would be easy to design astronomical devices to take advantage of this characteristic, perhaps by using small, symmetrical craters to support radio antennas or large mirrors. With no intervening atmosphere, telescopes operating on the lunar surface during the fourteen-day lunar nights might provide the best ‘‘seeing,’’ or ‘‘listening,’’ that astronomers could hope to find nearby in our solar system. We proposed to study such instruments for inclusion in the science payloads of these longer missions following the Apollo landings.

Compared with Apollo, where we were told there would be constraints on all the important exploration parameters such as payload weight, surface staytime, and site accessibility, we could think big. The biggest constraint to be removed was the limit on the payload we could send to the Moon’s surface. Instead of numbers like 250 pounds, we could plan around payloads of 7,000 pounds or more, which in turn could be used for any need we had. Experiments, life support, and transportation headed the list of items we would try to define so as to take advantage of the larger payloads.

As it was with Apollo, the astronauts’ safety was always uppermost in our thoughts as we laid these plans. Other self-imposed criteria required automat­ing as many jobs as possible to conserve the astronauts’ time. Lunar surface tasks would be designed to optimize their inherent ability to accomplish those aspects of exploration that humans do best: observing, describing, manipulat­ing complex equipment, and responding to the unexpected. We did not want them performing a lot of manual labor if it could be avoided. But we had to strike a delicate balance between automated functions and manual tasks, or supporters of unmanned exploration, both inside and outside NASA, would raise many questions and objections. Why go to the expense, not to mention risk, of sending astronauts if all they did was turn a switch and let a machine do the work? Switches could be turned on and off from Earth. Our office never thought this was a real challenge, since the astronauts’ unique abilities would always be their most important contribution toward exploring the Moon. A combination of automated equipment and hands-on tasks would be needed, and we took it for granted that exploration would proceed in this fashion.

Designing a drill for studying subsurface conditions (called logging) on the Moon and for taking subsurface core samples was a good example of how we eventually applied these criteria. On Earth these operations are labor intensive, requiring many types of laborers and technicians to carry out the wide variety of jobs each entails. Being familiar with all these tasks after spending many months at well sites in Colombia, I could see that new thinking would be required. Terrestrial drilling, logging, and coring equipment must be bulky and heavy to accommodate difficult drilling conditions and the constant rough handling encountered in the field.

Drilling on Earth has one other important characteristic that would be different on the Moon. Water or water-mud mixtures are normally pumped into a drill hole to cool the bit, bring the rock cuttings to the surface, and keep the hole from caving in. Where a water mixture cannot be used, air is circulated under high pressure to accomplish the same purposes. Either of these methods would be impractical on the Moon; we would have to find other ways. Since the primary purpose of drilling on the Moon would be to extract a core, we didn’t want astronauts to have to constantly oversee the drilling and coring. This added another dimension to whatever designs would be proposed: a highly reliable, semiautomated lunar core drill. We envisioned much more elegant equipment than that employed on Earth—probably to be used only once at each landing site and thus far different from traditional terrestrial designs.

With all these considerations to be dealt with, the next priority after we started the Martin study was to find a contractor who would do an overall analysis of science needs for the ALSS missions. This new study would generate first-order estimates of weights, volumes, and data transmission and power requirements for a suite of instruments selected by the government. This was my first attempt at writing a government request for quotation (RFQ), and I got help from my office and the NASA headquarters Procurement Office. The RFQ, called “Scientific Mission Support Study for ALSS,’’ focused on the scientific operations that could be done from a mobile laboratory carrying two astro­nauts. It was released in early 1964 from our headquarters office.

While I was writing this RFQ it became clear that managing contracts from headquarters would be difficult since we had so many studies to get under way. We needed to find a NASA center that would agree to manage them. Also, we reasoned that having a center take ownership of the studies had another advan­tage. The center would be a strong voice supporting our ideas at other NASA offices that might be skeptical of their importance when budget time rolled around and we were competing for scarce funds.

My few brief encounters with the MSC staff had not been encouraging. They were focused on Gemini and just beginning to think about Apollo science. As shown by our briefing to Faget, planning what should be done after Apollo was not on their agenda. In addition, in early 1964 I could not identify anyone I thought had the right background to manage the studies. Goddard Space Flight Center had built a strong earth sciences staff that could have taken on these studies, but they reported to the Office of Space Science and Applications, the wrong part of NASA. The Kennedy Space Center, although an OMSF center, did not seem to be an option, since its primary responsibility was to service a variety of launch vehicles and there were few earth scientists on the staff. That left the Marshall Space Flight Center, the remaining OMSF center, as my only choice. It turned out to be a most fortuitous final candidate. The studies initi­ated by our office and others in Advanced Manned Missions to improve the Apollo hardware had been undertaken by several MSFC organizations. Many MSFC staffers had worked on studies reported in the multivolume Lunar Logis­tic System.

Wernher von Braun, a German expatriate rocket genius, was the newly appointed MSFC director. He had just been reassigned from his position as director of the Development Operations Division of the Army Ballistic Missile Agency at the army’s Redstone Arsenal, located with MSFC in Huntsville, Ala­bama. At the end of World War II the army had brought more than 120 Ger­man engineers and scientists, led by von Braun, to the United States to improve the country’s rocket know-how. Some of this original group had been assigned to Cape Canaveral as well as Huntsville. With a perfect launch record for their rocket designs, they successfully launched the first United States satellite, and our rocket technology was progressing rapidly. Sending men to the Moon was to be their next challenge, which would include building the huge new Saturn V! MSFC was NASA’s largest center in terms of manpower, so the question became where to go in this organization, with which I had had no previous contact. The decision turned out to be easy, since the Research Projects Laboratory (RPL), under Ernst Stuhlinger, one of von Braun’s original team members, had been responsible for writing volume 10, Payloads, of the Lunar Logistic System re­port.3 This volume described science payloads that could be carried on modi­fied Apollo spacecraft, including many geophysical experiments.

After several phone calls I scheduled a meeting with James Downey, manager of the Special Projects Office in RPL; he and some of his staff had also contrib­uted to volume 10. Our first meeting took place in late 1963 and was marked by some careful bureaucratic dancing. Reflecting his center’s and his immediate boss’s cautious, Germanic approach to having someone from headquarters ask for a commitment of manpower and center resources, Jim wanted to know if my request represented a formal headquarters assignment of new duties for MSFC. I wasn’t prepared for such a pointed inquiry and knew I didn’t have the authority to say yes, so I hedged but assured him that our office had funds to support the studies I was asking him to manage.

Jim, a University of Alabama graduate, was an easygoing manager who commanded the respect of his unusual, multitalented conglomeration of scien­tists and engineers. He was eager to take on this new job, for so far his office had not received much funding for its studies. An important measure of a successful manager at NASA was how much funding he obtained and how many contracts he managed, so the promise of new funding was well received. But before he could agree it would have to be formally requested through the proper chan­nels. From my brief exposure to his staff, it appeared that they had the mix of skills needed to monitor the wide range of contractor studies we wanted to perform. I told Jim I would go back to Washington and start the paperwork. This meeting was the beginning of a long and productive relationship with Ernst Stuhlinger, Jim Downey, and their staffs as we undertook several studies that broke new ground for lunar exploration.

What did it mean when a NASA center managed programs or studies? There were many responsibilities. We met frequently to plan future procurements to be sure we all agreed on what the final products would be, and we would estimate the funds required and the schedules to be met by the contractors. Then MSFC would write the request for proposal (RFP), designate a contract monitor on Downey’s staff, establish a rather informal source selection com­mittee to evaluate the proposals, advertise the procurement in the Commerce Business Daily, release the RFP, evaluate the proposals received (with the evalua­tion documented in case of a protest from a rejected contractor), choose a win­ner or winners, award the contract, and then—the important part—monitor the contractor’s performance until the job was completed. The procedures we followed for these smaller contracts, although spelled out in NASA regulations, were nowhere near as precise as today’s requirements, which call for formally appointed source evaluation boards and source selection officials. Without this time-consuming bureaucratic red tape, we were able to move ahead quickly on our contracts.

In my mind the steps named above more than justified asking a center to help get the contracts under way; the centers had much more manpower avail­able for this cradle-to-grave job, as well as experience in directing the efforts of NASA’s growing number of contractors. The main responsibility of NASA headquarters staff was to develop the big-picture programs and run inter­ference with the administration and Congress on issues pertaining to budgets and policy, leaving the details of running the programs to the centers. In real­ity these distinctions weren’t so clear-cut, and the centers and headquarters worked together on all aspects of the programs. Contract management of advanced (paper) studies migrated more and more from headquarters to the centers. As NASA matured as an agency, the centers became powerful indepen­dent entities, supported by their homegrown political allies in Congress and the executive branch. This growing independence was one of the reasons friction developed between headquarters and MSC. Under von Braun, MSFC accepted headquarters direction more graciously; perhaps this smoother relationship was a reflection of MSFC’s confident corporate personality, embodied in the person of its director and enhanced by its established reputation in rocketry. MSC was the new kid on the block, attempting to prove that it knew how to get the job done but with a short track record. And it had no one with a reputation like von Braun’s to intervene if problems arose. Little by little, of course, MSC established this track record with the successful completion of the Mercury and Gemini programs, but this newfound confidence never translated to a smooth management relationship with our headquarters office in matters dealing with science.

Once MSFC agreed to manage our post-Apollo science studies, events moved rapidly. Contracts were signed in 1964 for the studies mentioned above, and soon afterward management of the ALSS Scientific Mission Support Study, won by the Bendix Aerospace Systems Division, was transferred to MSFC. Not all headquarters managers followed this practice; some liked to maintain con­trol of their programs by doing the day-to-day management. But the advan­tages of leaving contract management to MSFC were evident from the start. Small study contracts could be managed by headquarters staff, since they re­sulted only in paper, but once prototype hardware became deliverable, only a center could supply the management expertise and resources needed. Several of our contracts required delivery of engineering models or “breadboards” of proposed equipment as well as detailed analyses.

In June 1964, along with some reorganization at headquarters, the ALSS program was modified and given a new name, Apollo Extension System (AES). The new name was meant to convey a different message than Apollo Logistics Support System; AES was to be a new program based more closely on Apollo but not requiring the extensive hardware modifications envisioned for ALSS. There would still be a greater potential to study the Moon, both on the surface and from lunar orbit. We could still plan on dual launches of an automated LEM shelter-laboratory and a LEM taxi to carry the astronauts to the surface and return them to rendezvous with a CSM built for extended staytime. Our

strategy, as we had planned for ALSS, centered on the astronauts’ transferring to a shelter-laboratory after landing and conducting their extravehicular activities from there. AES studies also included using a wide variety of instruments aboard the Apollo CSM in Earth and lunar orbit to survey and map the surfaces of these two bodies. The orbital studies would now be managed in the Ad­vanced Manned Missions office as a continuation of the work initiated earlier by Pete Badgley.

In early 1964, President Johnson asked NASA to develop long-range goals for the agency and, by implication, the nation. Homer Newell, as was the custom, quickly asked the National Academy of Sciences to help provide a response focusing on space science. In 1961 the Academy’s Space Science Board (SSB) had recommended that “scientific exploration of the Moon and planets should be clearly stated as the ultimate objective of the U. S. space program for the foreseeable future.’’ Now, three years later, Harry Hess, chairman of the Space Science Board, wrote to Newell indicating that a change in objectives was appropriate. Planetary exploration, starting with unmanned exploration of Mars and eventually leading to manned exploration, should be the new goal.4 The SSB stated that Mars “offers the best possibility in our solar system for shedding light on extraterrestrial life.’’ It was ready to concede that the Apollo program would be successful, thus the new emphasis on planetary exploration. But the SSB also suggested some alternatives that included extensive manned lunar exploration leading to lunar bases. These recommendations, which we took as an endorsement of the studies we were pursuing, were eventually incor­porated into the report that was sent to the president. In the fall of 1964 we believed our programs would soon be officially embraced by the administra­tion, and this belief was reinforced a few months later when the president publicly declared that ‘‘we intend to not only land on the moon but to also explore the moon.’’5 We waited in vain for a formal start. Instead Johnson focused on his Great Society programs and, increasingly, on the war in Viet­nam. There were three more years of growing budgets for Manned Space Flight to fulfill the lunar landing mandate, but NASA’s overall funding peaked in FY 1965 and thereafter began to decline.

At the end of 1964 Ed Andrews and I were transferred from Tom Evans’s office to a new office called Special Studies under the direction of William Taylor. I was not pleased with this move; the mission of this new office was poorly defined, and it removed me from the day-to-day oversight of the pro­grams I had initiated. I maintained contact using my other hat, however, work­ing for Will Foster. Evans was promoted to lieutenant colonel that summer, and soon he left NASA and the army to return to Iowa and manage his family’s large farm. With his departure, the Advanced Manned Missions Lunar and Planetary Offices were combined under Frank Dixon, who until then had been director of the Manned Planetary Missions Office.

In June 1965 I was transferred back to Manned Lunar Missions Studies, once again a separate office, under a new director, Philip Culbertson, brought in from General Dynamics to replace Evans. I mention these office moves only to illustrate the uncertainty that was present at NASA as top management tried to position the agency for life after Apollo. Although Manned Space Flight’s bud­gets were still growing, management could foresee that if new missions were not assigned soon, the agency would be largely marking time until the end of Apollo. The mantra in OMSF was that only large, manned-mission programs could sustain NASA. Other programs, such as unmanned space science and aeronautics research, though important, would never maintain a prominent agency in the federal government’s hierarchy, which consists of large cabinet – level departments and also smaller independent agencies like NASA. In Wash­ington, big, growing government programs were good for those managing them, and declining budgets were bad for ambitious managers.

At the same time as we were attempting to define the science content of the ALSS-AES missions, the Boeing Company’s lunar base study, with the title Lunar Exploration Systems for Apollo (LESA), was under way. When William Henderson joined our office at the end of 1963 he became the headquarters lunar base expert and assumed oversight of all the lunar base studies. Boeing’s final LESA report described a modular lunar base that would be assembled from Apollo hardware, incorporating greater modifications than required for ALSS-AES missions. By grouping modules, a base could support colonies of two to eighteen men. (We had no women astronauts at that time, so the studies were always described in masculine terms.) Individual modules might take as much as 25,000 pounds of useful payload to the lunar surface. Depending on the mix of equipment and the number of modules, these colonies could operate for ninety days to two years. We envisioned sending to the Moon large pieces of scientific equipment that would permit a wide range of activities. Long – duration geological and geophysical traverses in large wheeled vehicles could be conducted, as well as studies confined to the base, such as deep drilling and astronomical observations. These endeavors, we believed, would lay the groundwork to justify permanent bases.

During this period we persuaded our management to let us take several trips overseas to gain greater insight into some of the situations we expected to encounter during lunar exploration. In January 1964 Bill Henderson took the first of such trips, receiving permission to visit our scientific bases in Antarctica. He made the case that these stations were the closest examples we could find to what a base on the Moon would be like: isolated, difficult to supply, and therefore self-sufficient. Their primary reason for existence was to conduct scientific investigations; the secondary objective was to show the flag—or per­haps vice versa. Both these reasons closely followed what we believed would be the ultimate rationale for establishing lunar bases, and one couldn’t deny that Antarctic conditions were moonlike. Bill thought his time in Antarctica was well spent and, since he was the only person at headquarters with this ex­perience, his recommendations carried more weight when he advanced his thoughts on how to design a lunar base.

At the end of the rather massive Boeing study, Bill initiated a new round of more detailed lunar base analyses. The resulting contract, signed by the Lock­heed Missile and Space Company in February 1966 for $897,000, was the largest award ever made by our office. The study, called Mission Modes and Systems Analysis, would be supported by three other contractor studies valued at an additional $900,000. One of these studies, Scientific Mission Support Study for Extended Lunar Exploration, was won by North American Aviation, with Jack Green, of the ‘‘volcanic Moon,’’ playing a prominent role in the study. The contract would be monitored by Paul Lowman and Herman Gierow, Jim Downey’s deputy and a versatile manager who had participated in the earlier LESA studies.

For decades space dreamers and enthusiasts, including MSFC’s director, von Braun, had written and lectured on the possibility of establishing a lunar base. Now major government funds were to be spent on a serious look at what it would take to carry it off. The inherent ability of the Apollo hardware to place large payloads into Earth orbit and send them on to the Moon was the initial requirement for lunar base planners. After modifications, with each flight the Apollo upper stages would be capable of placing large payloads on the lunar surface. Big payloads meant you could envision supporting and supplying a large lunar colony over long periods at a reasonable cost. This was the challenge, first to Boeing, then to Lockheed and its support contractors: Tell us how it could be done, what such a base would look like, and how a base could support scientific and engineering operations that would justify its existence. The results of all these studies were encouraging, especially assuming that the nation would continue to commit large amounts of money to the investment it was making in Apollo—not an unreasonable expectation in the mid-1960s. Extended lunar exploration, followed by the establishment of one or more lunar bases, would not be cheap. But the initial analyses seemed to show that, for an additional investment approaching what would be spent on Apollo, all this could be done.

Bob Seamans, George Mueller, and E. Z. Gray began to lobby Congress for a NASA mandate that would implement these grand designs. When they testi­fied before NASA congressional oversight committees, they would impress the members with realistic artists’ renditions of what these stations and bases could look like. They also had funding estimates (supplied from our contractor stud­ies) to support their contention that continued lunar operations were feasible at a reasonable price and would produce important results. At a lower level in the management chain, staff like me, Paul Lowman, Bill Henderson, and others involved in the studies at MSFC took every opportunity to advertise our plans at professional conferences and public forums. We could usually count on good coverage from the media, and it seemed at the time that we were winning public support. Public polls always gave NASA high marks, and the major news and trade magazines were eager to write stories and show drawings of future lunar colonies.

Contractors who won our awards usually included well-known scientists on their teams as consultants (a few with Nobel credentials); they were to review study results during the contract and make recommendations to the contrac­tors to ensure that the results were grounded in scientific reality. During pro­posal evaluations, the quality of these consultants could determine which con­tractor would receive the award. While the contract was under way, or at its conclusion, we were not bashful about dropping their names if our assump­tions were challenged.

Returning to the ALSS-AES studies, in May 1964 MSFC put together the RFP for what we called the Emplaced Scientific Station (ESS). This study would provide a preliminary design of a self-sufficient geophysical station to be de­ployed by the astronauts on the lunar surface, incorporating several experi­ments listed in the Sonett Report and some from other sources. We received eight responses to the RFP and selected two contractors, Bendix Corporation, led by Lyle Tiffany, and Westinghouse, led by Jack Wild. These two contracts, along with the Scientific Mission Support Study, would provide us with enough detail that one year later we could extrapolate the results to design the Apollo geophysical station, which would have to meet more stringent requirements.

As we did for the ESS, we awarded two contracts in 1965 to study competing designs for a hundred-foot drill. One went to Westinghouse Electric Corpora­tion and a second to Northrup Space Laboratories. Each contract had a value of more than $500,000. The MSFC contract manager was John Bensko, a geologist who had worked in the oil and coal mining industries before joining NASA. After coming to MSFC, he helped develop engineering models of the lunar surface, useful background for his drill contracts. John put together an advisory team from the Corps of Engineers and the Bureau of Mines to provide addi­tional engineering expertise as the contractors began to cope with their difficult assignments. In those days NASA always attempted to at least match the con­tractors’ expertise in house so that our oversight and evaluation of their perfor­mance were well grounded. I believe this respect for each other’s abilities let NASA and its contractors work together better as a team, although some con­tractors grumbled at the tight monitoring. Today NASA’s approach to contract monitoring seems to have changed almost 180 degrees; in-house expertise in the aspects of a contract is often minimal. For the drill studies, NASA’s compe­tence was especially important, since we planned a series of difficult tests in­cluding drilling in a vacuum chamber at MSFC, never before attempted with a drill of this size.

Considering the unusual location for a drill rig and other constraints, the Westinghouse approach to drilling on the Moon was relatively straightforward, modeled after terrestrial wire-line drilling. Short sections of drill pipe were added from a rotating dispenser as drilling progressed; the core would be extracted from a short core stem after each section was taken from the drill hole. Since this would be close to a conventional design, it would entail almost constant monitoring by the astronauts. The Northrup design was radically different. It proposed using a flexible drill string, wound on a drum, that would be slowly fed into the hole to the final target depth of one hundred feet. A core stem would be attached to the end of a flexible pipe, and the core would be recovered much as in the Westinghouse design but without adding drill pipe sections every five to ten feet. Several innovative concepts were aimed at reduc­ing the astronauts’ involvement, and though we recognized that they posed some design risks, we accepted them as the price for a possible breakthrough in technology.

One of the major challenges for both concepts was cooling the bit during drilling to reduce wear. Bensko hired Arthur D. Little to do a separate analysis of how to accomplish the cooling. The company’s study showed that the cool­ing problem could be greatly mitigated in the vacuum environment of the Moon if the rock cuttings could be rapidly moved away from the bit face so that the they would carry off some of the heat. Spiral flutes were thus incorporated on the outside of the drill string, like an auger, to lift the cuttings up through the hole to the surface.

Although the spiral flutes partially solved how to cool the bit, as our studies progressed we found that after a short time the bit would still get too hot, become dull, and stop cutting. Both contractors settled on using diamond-core bits to ensure that they could drill through any rock type encountered. Westing – house had included Longyear on its team, and Northrup had teamed with Christianson Diamond Bits, the leading industrial suppliers of diamond-core bits. Both bit contractors concluded that, with the technology then available, even a diamond-core bit would need to be replaced many times in drilling a hundred-foot hole. This was unacceptable.

Initially, the best the Westinghouse team could do under test conditions was to drill fourteen inches through basalt, a possible lunar rock type, before an uncooled bit failed. But they reexamined the problem and finally hit on a solution. The diamond-core bits then offered to industry used a matrix that ‘‘glued’’ tiny diamonds to the bit in a random alignment. The random align­ment did not allow each diamond to present its best cutting edge to the rock being cored, however. They demonstrated that carefully setting the diamonds in the matrix significantly prolonged the life of the bit. Hand setting each diamond would add greatly to the bit’s cost, but it would be well worth it for a lunar mission where the astronauts’ time was more precious than a diamond bit. These newly designed bits lasted more than ten feet before they dulled. After other design changes, eventually we expected to drill the entire one hundred feet with just one bit, eliminating a time-consuming chore. As I recall, Chris­tianson developed a relatively inexpensive technique to manufacture bits of this design for their terrestrial customers. Although they cost more than normal diamond-core bits, they were worth the investment because fewer were needed.

The cost of drilling on Earth is strongly influenced not only by the price of bits but by the time needed to extract a dulled bit from the drill hole, change bits, and resume drilling.

As the studies continued, progress on the Northrup design slowed, and the contract was terminated before they delivered a complete working model. Our gamble had failed. A Westinghouse model was tested at MSFC, including vac­uum chamber tests. Finally tests were held in the desert in Arizona and New Mexico to simulate drilling under lunar conditions (but not in a vacuum), with no lubrication for the bit. Bensko recalls that we chose a bad time for our tests: there had been more rainfall than normal, and the wet soil gummed up the flutes. In other tests the fluted drill pipe performed about as expected, and we were encouraged to believe that a full-scale drill could extract cores on the Moon to depths of one hundred feet.

In anticipation of drilling a deep hole on the Moon, in 1965 we started two studies with Texaco and Schlumberger to design logging devices that would determine conditions beneath the lunar surface. (Taking measurements in ter­restrial drill holes is standard practice for obtaining information on subsurface conditions.) These contracts, also worth more than $500,000 each, were man­aged by MSFC’s Orlo Hudson.

In both terrestrial drilling and drill-hole logging, the drill hole is almost always filled with a fluid, of varying chemistry, the remnants of the drilling mud. Lacking this liquid to couple the logging tools to the subsurface rock formations, the contractors were forced to modify standard oil field technology. The Texaco team, which had extensive experience in developing logging devices for oil field exploration, had won an award from the Jet Propulsion Laboratory (JPL) to provide logging devices for the Ranger and Surveyor projects. In their planning stages both projects included small drills as potential science pay­loads. Schlumberger, the acknowledged leader in developing logging devices for the oil and mineral exploration industry, showed an interest in such unworldly studies (to our surprise), entered a bid, and won the other contract. Both contractors overcame the lunar logging constraints and designed a suite of devices that could make measurements in a hole drilled on the Moon. Perhaps one day, when the opportunity arises to drill deep holes on the Moon or some other extraterrestrial body, these studies will be found and reread.

The most interesting set of studies we conducted were those related to providing mobility once the astronauts reached the lunar surface. Many con­cepts were being proposed, some more fanciful than others. MSFC had re­ported the results of the first in-house mobility studies in volume 9 of the Lunar Logistic System series.6 Two of the main contributors to these studies were Jean Olivier and David Cramblit, who wrote several reports on lunar surface mobility. To learn what types of mobility systems would work best on the Moon, based on the limited knowledge available, MSFC and the Kennedy Space Center developed a lunar surface model to study how wheeled vehicles might perform on soils in a lunar vacuum and what type of obstacles they would have to traverse.7

JPL had also developed a lunar surface model in order to design a small unmanned vehicle for the Surveyor project.8 It had tested several designs on simulated lunar terrain in the early 1960s. My first trip to JPL was to witness a test of a small vehicle operated by an engineer with a handheld remote-control box, hardwired to the rover. It was much like a modern toy car except for the connecting wire. Today’s electronics permit cheap radio-controlled toys; in the early 1960s radio control was a luxury we usually did without when testing our concepts. This was an interesting demonstration of a small articulated vehicle with springy wheels driving over loose sandy material and small rocks. From time to time there were short interruptions caused by failures in the then state – of-the-art electrical circuits, powered by vacuum tubes. One could say that the granddaughter of this vehicle was the small rover named Sojourner that tra­versed the Martian surface in July 1997. A United States automated rover never made it to the Moon, but a Soviet rover named Lunokhod operated on the Moon in 1970.

Although in 1964 and 1965 we still did not have any data from direct contact with the lunar surface, information from radar and laboratory studies pre­dicted how the Moon’s surface layer would respond to a wheeled vehicle. In spite of Tommy Gold’s theories, we were certain that a vehicle could move around without serious difficulties. But we were not sure how the Moon’s almost total vacuum would affect the lunar soil; the high vacuum that would be encountered on the Moon was impossible to achieve on Earth. Studies had been conducted in high vacuum using several types of simulated lunar soil, but their fidelity was open to question because our ideas about the composition of lunar soil (grain size, mineralogy, and other characteristics) were mostly guesses.

Our first contractor studies of a lunar surface vehicle were undertaken by the Bendix Corporation and the Boeing Aerospace Division. They were selected in

May 1964 to study ALSS exploration payloads, including a vehicle we had dubbed MOLAB (for mobile laboratory). The Boeing study was managed by Grady Mitchum, and the Bendix manager was Charles Weatherred. Because of their involvement in the post-Apollo studies, both these men and their com­panies would be important contributors to later Apollo contracts. Bendix had earlier won one of the JPL design contracts for a small Surveyor rover, so it was well prepared to undertake the study. From taking part in our lunar base studies, Boeing had a good background that included designing mobility concepts.

The concept for using a MOLAB was to have it delivered to the Moon by an ALSS automated LEM. It would then be deployed and operated remotely so that it could travel to another LEM carrying two astronauts that would land a short distance away. It was to be a vehicle of about seven thousand pounds, including the scientific equipment it would carry. It would support two astro­nauts for up to two weeks in a pressurized cab, permitting shirt-sleeve working conditions while under way. Based on our study of early geologic maps of the Moon, we felt that such a vehicle should have a traverse range of several hun­dred miles so the astronauts could make several trips far enough from their landing site to sample geologically interesting areas. These requirements were a tall order for any vehicle, not to mention one that must function on the lunar surface.

The two contractors were also asked to design a shelter that could be deliv­ered by the same type of automated LEM and a smaller, unpressurized vehicle we named the local scientific survey module (LSSM). (Moon vehicles had to have strange names; they couldn’t just be called cars or trucks, since they would be so different from any of their terrestrial cousins.) All these studies were to be accomplished by both contractors for a total of slightly more than $1.5 million.

As the studies progressed, under the direction of Joe de Fries and Lynn Bradford at MSFC, the MSFC Manufacturing Engineering Lab built a full-scale mock-up to evaluate such things as cabin size and crew station layout. Many photographs of this rather unusual looking vehicle were circulated to the media and other interested groups, showing our progress toward the next step in lunar exploration. A December 1964 issue of Aviation Week and Space Technology featured a front cover picture showing the mock-up sitting on top of a LEM truck and included a special report on the Bendix version.9 The MOLAB, more than any other project we worked on for post-Apollo missions, seemed to catch the imagination of futurists, perhaps reflecting the national love affair with the automobile. Perhaps people could visualize themselves speeding across the lunar surface, dodging boulders and craters.

At the conclusion of the initial contracts in July 1965, both contractors were given extensions totaling more than $1 million to refine their LSSM designs. Bendix and General Motors received two other contracts to produce four-wheel and six-wheel LSSM test designs, each worth almost $400,000. By the end of 1965 we had awarded lunar vehicle contracts for more than $3.5 million and had probably spent almost as much for in-house civil service workers and contractor support.

While all this wheeled-vehicle planning was under way, Textron Bell Aero­space Company was quietly developing a small manned lunar flying vehicle (LFV). A one-man version was demonstrated in a live test early in 1964. (A later generation of this device was demonstrated at large gatherings including the 1984 Olympics in Los Angeles, and a version was flown in the James Bond movie Thunderball.) Bell had conducted a preliminary study of how to com­bine the MOLAB and the LFV, sponsored by NASA’s Office of Advanced Re­search and Technology. In these early days we had a good working relationship with OART; under the direction of James Gangler, it was attempting to look far ahead at technology needs for lunar exploration and lunar bases. After the impressive one-man flight demonstration, MSFC awarded Textron Bell a follow-on contract in August 1964 to further define the concept. In these stud­ies the LFV was given two functions—to return the astronauts to a base camp in case of a MOLAB breakdown and to help them reach difficult sites.

The MSFC contract with Textron Bell called for an LFV design that would carry two astronauts a minimum of fifty miles for the safety fly-back mission. This would also be a useful range to take the astronauts to sites they could not reach overland. MSFC later awarded Bell a second contract with a more modest goal—to support AES missions requiring an operations radius of only fifteen miles. This vehicle, which needed far less fuel because of its shorter range, could carry one astronaut and three hundred pounds of equipment or transport two astronauts the same distance. Both design studies and a working prototype indicated that an LFV with these characteristics was feasible.

A study was also done to assess the advantages of using the lunar surface for astronomical observations, an application supported by some, but not all, in the astronomical fraternity. In 1965 MSFC awarded Kollsman Instrument Cor­poration a one-year contract for $144,000 to assess the feasibility of carrying a large optical telescope observatory to the Moon mounted on a modified auto­mated LEM lander. MSFC’s contract monitor was Ernest Wells, an amateur astronomer whose avocation served him well in this job. Kollsman was already developing the Goddard Experimental Package (GEP), an automated observa­tory scheduled to be launched in 1966 on the Orbiting Astronomical Observa­tory (OAO), so working with the company would save effort and money.

The GEP consisted of a thirty-six-inch reflector telescope, its mounting, a camera, and associated electronics. Improvements to the GEP design to take advantage of its lunar location could be recommended during this study, as well as design changes to accommodate the astronauts’ involvement in its operation, since the OAO design was a fully automated observatory. The results were encouraging, indicating that the astronomical payload could operate on the Moon for long periods in both an unmanned and a manned mode.10 Kollsman also reported that new technology, by greatly reducing the overall weight, might permit a much larger instrument, perhaps up to 120 inches in diameter, to be carried on the same LEM truck.

A fallout of these studies at MSFC was the establishment of a Scientific Payloads Division in Stuhlinger’s Space Sciences Laboratory. Jim Downey be­came the director of this new division, and Herman Gierow was named deputy. Later, as the MSFC work on post-Apollo science wound down, both Jim and Herman went on to manage important new programs that included work on the Apollo telescope mount flown on Skylab. Their work on space-based astronomy culminated in the launch of three high energy astronomical obser­vatories in the 1970s and studies of a large space telescope that evolved a few years later into the successful Hubbell space telescope program.

The transition from planning ALSS missions to planning AES missions was relatively painless. AES payloads would be smaller than those we anticipated for ALSS missions but much larger than Apollo’s allocation. By this time we had a much better understanding of the Apollo hardware than when we started our ALSS studies, and we were also becoming aware of the potential Apollo opera­tional margins that could permit larger payloads or increase flexibility. We hoped these margins would soon be available as confidence in Apollo’s perfor­mance grew.

Removing the ascent propulsion and other unnecessary systems required during a normal LEM ascent and rendezvous would free up space for approxi­mately 6,000 pounds of payload, 1,000 pounds less than the total used for the

ALSS studies. Of the 6,000 pounds, 3,500 would be required for consumables and other additions so two men could stay in the LEM for two weeks. The remaining 2,500 pounds could then be used for scientific equipment. This represented a rather firm increase of an order of magnitude over the expected allocation for Apollo science payloads. Although 2,500 pounds was less than half the weight we had been using in planning, it was enough to be exciting.

Based on 2,500 pounds and results coming in from our ALSS-AES studies and USGS work at Flagstaff, we divided a typical payload as follows: 1,000 pounds for a fully charged LSSM with a range of 125 miles, 200 pounds for a hundred-foot core drill, 90 pounds for logging devices, 350-400 pounds for an ESS, 80 pounds for a small preliminary sample analysis lab, 100 pounds for geological field mapping equipment, 150 pounds for geophysical field survey equipment, 30 pounds for sample return containers, and up to 500 pounds for a power supply for the drill or other exploration equipment. We felt this equip­ment would let the astronauts take full advantage of a two-week stay and study their landing site in some detail. For safety reasons, during manned operations the LSSM would be restricted to a radius of five miles, but it could operate in both manned and automated modes. After the astronauts left it could carry out investigations farther from the landing site, to the limit of its battery charge, under command from Earth.

Our planning for lunar exploration after the initial Apollo landings was now in high gear. The next step was to test our ideas as realistically as possible so we could not be accused of offering proposals thought up by ‘‘some high – school student.’’

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.