Developing the Geological Equipment,. Related Experiments, and Sampling Protocols

Methods of conducting geological field studies have changed little in the past two hundred years. The geologist visits the locale to be studied, samples rocks, measures structural features like hills, valleys, cliffs, and other surface topogra­phy, traces formation boundaries (if possible), determines the relative ages of these various features, usually by several techniques, then interprets this infor­mation and finally makes a map. Aerial and satellite photos, as well as new surveying instruments and global positioning systems, now simplify and speed up the fieldwork, but all these steps are still necessary to produce a final map. In many cases geophysical data can help in making subsurface interpretations, but the overall job remains the same: sample, measure, interpret. Depending on the geological complexity of the site and the geologist’s skills, this can be a time­consuming endeavor. Some sites have been studied for years by the same or different geologists, slowly yielding an interpretation that most workers will agree with.

Lunar geological fieldwork would present the same challenges that faced a terrestrial geologist plus many more. For example, at the beginning of Project Apollo it was not clear how easily astronauts could sample and measure lunar features; above all, in spite of the many hours spent in geology training, it was questionable how skilled they would be at deciding how and where to sample and take measurements. Each Apollo landing site would represent a one-shot opportunity to collect as much information as possible—there would probably be no return to resample or remeasure—so it had to be done right. This de­mand haunted the new breed of ‘‘lunar geologists”: they had to complete the job the first time. That very little hard data would be in hand until the Apollo landings took place (Ranger, Surveyor, Lunar Orbiter, and ground-based obser­vations notwithstanding) added enormous complications for those of us at­tempting to prepare the equipment that would be taken on each mission and to plan the exploration strategy.

In February 1964 Will Foster sent a set of recommended Apollo investiga­tions and investigators to the Space Science Steering Committee (SSSC),1 the group Homer Newell had charged with advising him about what science to conduct on all space programs. In his memo Foster listed five areas of Apollo investigations—geology, geochemistry, geophysics, biology, and lunar atmo­sphere—and named scientists who should be on the investigating teams. As expected, the recommended geology fieldwork team was headed by Gene Shoe­maker. It included Hoover Mackin from the University of Texas, Aaron Waters from the University of California, Santa Barbara, and Edward Goddard from the University of Michigan. The geochemistry planning panel included James Arnold from the University of California, San Diego, Paul Gast, then at the University of Minnesota, Brian Mason from the American Museum of National History, and several other noted geochemists. Related to the geochemistry panel was the petrography and mineralogy team composed of Harry Hess of Princeton, Clifford Frondel of Harvard, Bill Pecora and Ed Chao of the United States Geological Survey, and Edward Cameron of the University of Wisconsin.

Shoemaker’s Field Geology Team was responsible for planning the lunar fieldwork, determining the requirements for maps and tools, monitoring the astronauts’ training and their activities once they reached the Moon, and pre­paring the necessary reports. Working with the geochemistry planning panel and the petrography and mineralogy team, the Field Geology Team would plan sample collecting procedures and design sampling equipment that would sat­isfy the needs of future sample-analysis PIs. For samples that would be returned to Earth, the geochemistry planning panel and the petrography and mineralogy team would recommend the protocols for sample preparation. Finally, the geochemistry planning panel was asked to recommend to Foster’s office par­ticular investigations and investigators for studying the samples. These teams and panels were subsequently approved by the SSSC and began their work.

Before Shoemaker’s appointment, two conflicting concepts for field geology instrumentation were under development, one designed by the staff at the Manned Spacecraft Center and the other by USGS in Flagstaff. MSC, led by Uel Clanton, had devised an engineering model of an all-in-one geological tool that the astronauts could use for sampling, drilling, and several other functions, in an attempt to simplify the many tasks they would have to accomplish and at the same time save weight and time by reducing the number of tools needed.

USGS had similar concerns but thought the biggest problem would be locating and documenting the sites visited, and in particular sampled, so that accurate traverse maps and profiles could be reconstructed back on Earth. The Flagstaff team had devised a surveying staff that would reflect a laser beam from a ranging device and automatically record the coordinates of a position on the lunar surface. This approach was based on the simulations and exercises we had been conducting for the post-Apollo missions, which suggested that without some type of surveying instrument it would be almost impossible for an astro­naut to accurately locate his position on the Moon and associate a sample or ob­servation with a specific point. Lunar geologic maps made without such posi­tioning would be seriously degraded in value, since to establish map locations we would have to depend on some type of dead reckoning or coarse Earth­tracking and reconstruction of the traverse based on voice communication.2

Our experience during the Martin Marietta contract, and the growing con­cern about measuring distances on the lunar surface, led the Branch of Astro – geology to further explore including a periscope in the lunar module (LM), as we had proposed earlier, rather than the sextant that was being planned for navigation. In February 1965 Gordon Swann and Dave Dodgen visited two navy periscope suppliers, Kollmorgan and Kollsman Instruments, to discuss their ideas. Besides the concerns arising from the Martin contract, they wanted to be able to track an astronaut if only one was allowed to leave the LM. Though both companies thought the Apollo navigation requirements and the surveying ability needed on the Moon’s surface could be incorporated in one instrument,3 no official action was taken. A jury-rigged optical ranging periscope built by David Dodgen and Walt Fahey was used during some field simulations to assess the value of such an instrument.

These three pieces of equipment had their advocates and their detractors. At the end of 1965 the MSC engineering model was tested by a joint review team composed of members of Foster’s office and several MSC offices, including representatives from the astronaut office, and we agreed to stop work on this tool. Because of its several functions, it was large and cumbersome, with so many batteries, handles, switches, and other components that it looked like a Rube Goldberg contraption. The USGS surveying staff survived our initial evaluations. In spite of the advertised versatility of these tools, the astro­nauts would still need additional equipment for tasks that the all-in-one de­signs could not perform. Converting the LM sextant to a periscope was also finally abandoned because of the added cost and schedule delay entailed by modifying the LM navigation system. For the last three missions, a navigation system on the astronauts’ lunar rover met most of the tracking and mapping requirements.

As we began to design and build prototype tools, another complication arose: certain materials and designs might interact dangerously with the space­craft’s atmosphere, communications, or even the astronauts’ space suits. These restrictions, some certainly necessary, would be a bone of contention through­out the equipment development phase, adding trouble and expense to what could have been, in some cases, rather straightforward procurements.

Without question, the most important task the astronauts would perform on the lunar surface would be sample collection. There was much debate on how best to do this. How much sample? What types of samples? How should they be packaged for the trip home? How badly would the lunar surface, and in turn the samples, be contaminated by the effluents from the LM descent engine plume? These questions and many more faced us as we began to realize that a lunar landing was not far off. The danger of contaminating the Earth was being addressed, but designing the sample containers to minimize this concern still lay in the future. Answers to all these questions would affect the design not only of the sample containers but also of the collecting tools.

To start answering the sampling questions, the Office of Space Science and Applications asked USGS to detail to NASA a person with experience in sample collection and analysis. Ed Chao was the first to arrive, soon followed by Verl Richard Wilmarth, a senior USGS manager. Dick arrived at NASA in early 1964, and I first met him soon afterward in his new office in federal office building 6. NASA shared FOB-6 at that time with other government agencies, and though it was older than FOB-10, where my office was, the building was more luxurious; wider corridors, bigger elevators, a fancier cafeteria, and the other trappings of power so important in Washington. The NASA administra­tor and senior staff had offices in this building as well as OSSA, the General Council, Legislative Affairs, Public Affairs, and several other NASA depart­ments. The top floors had been taken over by NASA, and some offices afforded a wonderful view of the city. The administrator’s office faced west toward the

White House, and Legislative Affairs looked east toward Capitol Hill—perhaps by some logic, though probably just by chance.

Although he was an experienced manager, Wilmarth had never had an assignment quite like this: soliciting the scientists of the world to bid for a piece of the returned lunar samples and perhaps a chance to win a Nobel Prize—a once in a lifetime opportunity. I told Dick about my experience in developing this type of solicitation, officially called an announcement of flight oppor­tunities (AFO), as well as my background in writing government requests for proposals (RFPs) that had been released from NASA headquarters. Lacking this experience, especially with the quirks of NASA procurements, he asked me to assist him in his new job.

For the next several months Dick wrestled with his task, and I spent a significant part of my time helping him. Many meetings and consultations with interested parties were needed to be sure we were not overlooking some large or small detail. The AFO had to ask for information covering several areas, in a form that would let a blue-ribbon panel, still to be identified, select the most qualified proposals. What was the objective of the analysis? How much sample was needed? Would the analysis involve destructive or nondestructive testing? What were the packaging requirements? What type of equipment would be used? Would there be collaborators in addition to the principal investigator (PI), and who would they be? How much funding would be needed? How long would it take to do the analysis? Finally, after several months of labor, a draft of the AFO was ready to be circulated to senior management, and after review by both OSSA and the Office of Manned Space Flight, a final version was released at the end of 1964. The AFO asked that proposals be delivered to NASA by June 1965.

Before the sample proposals were received, Shoemaker’s Field Geology Team began developing concepts for tools that could collect a variety of lunar samples as well as take the measurements needed to conduct geological studies. These designs were based on both the Sonett Report and the Falmouth conference report, with the latter providing some specific recommendations: a long- handled trowel (really a small shovel); a rock hammer; sampling tubes to be hammered into the lunar soil to collect small subsurface samples; a hand-held magnifying glass; a combination scriber and brush to mark and clean the samples; and sample bags and special sample containers, one of them airtight. A camera was also recommended. We began to build prototypes of these tools at

MSC and at Flagstaff, believing that eventually, regardless of whatever unique requirements we ultimately received from the still to be selected sample PIs, all these tools would be needed.

With the possible exception of the airtight container, these early tool and sample container lists constituted the standard inventory that any field geologist would recognize, modified for their unique application. Everyone knew, for example, what a geologist’s hammer looked like. But some changes would be needed, since each tool would be used by a space-suited astronaut, perhaps under difficult lighting and temperature conditions, and in one-sixth gravity. We also had to factor in limited payload weight and stowage space, both on the trip to the Moon and returning. We knew that meeting all these constraints would require some compromises, clever design, and perhaps most important, careful input from the astronauts.

In September 1965, shortly after the Falmouth conference, Will Foster sent MSC a proposed second set of guidelines for Apollo science. In his memo he asked Robert Gilruth, MSC center director, and Max Faget to ‘‘prepare a Pro­gram Plan from which we can establish firm Program Guidelines to which all of us involved in this effort can work.’’4 Foster’s guidelines included discussions of sample return and lunar sample boxes, the Lunar Receiving Laboratory (LRL), the geophysical ground station, recently given the name Apollo Lunar Surface Experiments Package (ALSEP), and the geological hand tools and other equip­ment. He urged MSC to develop the guidelines as soon as possible, since we had little time to deliver the scientific equipment for the first missions.

While these guidelines were being developed we continued selecting the sample analysis PIs. After their proposals were received, Dick Wilmarth, Ed Chao, and Bob Bryson spent the next several months visiting the potential PIs and their labs to determine if they were equipped to conduct the analyses they proposed. Some were, some were not. As a result, OSSA began a program to upgrade the labs even though their proposals had not been officially approved. During the next five years, NASA transferred over $19 million to the sample PIs to purchase equipment and compensate them for their efforts.

As part of its responsibilities, the Field Geology Team began a careful review of the proposals by establishing a geology working group chaired by Shoe­maker. In addition to Shoemaker, the working group consisted of Goddard, Mackin, and Waters from the Field Geology Team, Harry Hess (from the Space Science Board), and Ted Foss and Jack Schmitt from MSC. I served as secretary.

We met over a period of nine months, and at the end of 1966 we sent our report to OSSA. We recommended that almost all the proposals submitted be ac­cepted, a total of forty-one.5 At Dick Wilmarth’s urging we also submitted a list of tests and experiments that should be conducted at the LRL, the equipment the lab should contain, and based on our ongoing studies, the types of con­tainers that should be carried on the missions to hold the different types of samples we expected would be collected.

With Walter Cunningham immersed in his duties with Gemini and Apollo, our astronaut contact for the development of science equipment became Don Lind. Don had been selected in April 1966 as one of the nineteen astronauts in the fifth selection group, less than a year after the first scientist-astronaut selec­tion. He had a Ph. D. in physics, and I had worked with him at Goddard Space Flight Center, where he was employed before his selection. He was an excellent choice to interact with the science community. Since he had also been a navy pilot and had a reputation at MSC as a meticulous worker, his opinions carried a lot of weight with the astronaut office. Jack Schmitt, as the only geologist – astronaut, would become closely involved in designing and developing the tools and experiments, but at this time he was just finishing his flight training.

Lind became our sounding board and made important contributions to Apollo science. He spent many hours trying each new design in a pressure suit, and along with Gordon Swann and other MSC and USGS staff he attempted to validate them in NASA’s converted Air Force KC-135 (nicknamed the ‘‘Vomit Comet’’ for the reaction of many test subjects during the flight parabolas spe­cially calculated to provide short periods of low or zero gravity). Ray Zedekar and others from the MSC Flight Crew Systems Division also worked tirelessly to test and improve the tools.

Simulations continued at Flagstaff through 1966 and 1967, prompting con­siderable refinement in the number and design of the hand tools the Field Geology Team would recommend. Astronaut mobility, dexterity, and visibility in the pressure suit were ultimately the major considerations and led to several unique tools not carried by geologists on Earth. In February 1967 a critical design review (CDR) of the Apollo lunar hand tools was held at MSC.6 Because several of the proposed hand tools were not ready for the review, it was decided to designate a ‘‘hand tool pool.’’ From the pool, a total of about twenty pounds of equipment could be selected for each mission, tailored to the mission’s specific needs. A tentative priority list was established: tool carrier, sample bags (100-200), maps, tongs, hammer, scoop, drive tube number 1, extension han­dle (used with several tools to eliminate bending over), gnomon, drive tube number 2, surveying staff (later dropped from the pool), color chart, drive tube number 3, sample bag dispenser and sealer, aseptic sampler, spring scale, and combination brush/scriber/hand lens.

The tool carrier, a three-legged stand, allowed the astronauts to carry their tools from station to station with one hand and then reach them without stooping. It was used on only two missions, Apollos 12 and 14. A second design carried on the J missions held the tools so that they could be mounted on the rear of the lunar rover.

The gnomon, a unique device, was devised by USGS to be placed in the field of view of the cameras the astronauts used on the lunar surface. It provided geometric and photometric control so that the photographs could be used to make analytical measurements. It consisted of a tripod about fourteen inches high supporting a gimbaled, weighted rod that would hang vertically. The shadow cast by the rod (hence gnomon) showed the direction the camera was pointed so that the astronaut need not estimate it and transmit it by voice. A gray scale on the rod was used for photometric calibration of the black- and-white photos, and a color chart on one leg helped us calibrate the color photos. With all this data available, we were eventually able to make stereo pairs from the photos and produce contour maps of the areas where the photos were taken.

The spring scale would weigh the rock boxes and individual sample bags brought back to Earth. These weights were important to the engineers doing trajectory analysis during the astronauts’ return journey. Those who saw the movie Apollo 13 may remember that Mission Control in Houston could not understand why the returning spacecraft did not respond as expected to the course corrections being made to bring the astronauts back within the narrow corridor in space required for a safe reentry. The combined LM and command module (CM) weights were accurately known, so they should have responded predictably to the small thruster burns. Finally someone remembered that the computer programs had been calculated allowing for a few hundred pounds of returned lunar samples. No samples were on board, since the astronauts had never landed on the Moon. When this figure was corrected and the proper weight inserted into the programs, the returning spacecraft was steered pre­

cisely into the Earth’s atmosphere, allowing the command module to make a safe landing.

At this CDR, concerns again surfaced about the materials used in the tools. One dealt with the magnetometer experiment that would be deployed with the ALSEP and stowed near the tools on the LM. Stainless steel (the preferred material for the hammer and drive tubes, for example) might induce too much remnant magnetism, thus affecting the accuracy of its readings. Another con­cern was how hot or cold the tools would become in full sunlight or shadow, since the gloves used for extravehicular activity (EVA) could tolerate tempera­tures only in the range of —250°F to 175°F. It was decided that the tools would be anodized or given a gold tone to moderate temperatures on the surfaces the astronauts would touch.

Also at this CDR the surveying staff received a careful reexamination. To take full advantage of its capabilities the astronauts would have to make twelve settings at each station, taking a total of five to ten minutes. We were told the astronauts thought this was too long, and most of us agreed; their time on the lunar surface would be our most precious resource. The staff was eventually dropped from the pool. By the time the J missions flew, the ‘‘hand tool pool’’ was no longer required because the science payload was large enough to accom­modate all the needed tools, some of which were new to the J missions or had been redesigned by that time.

With this background, we can now turn to sampling. The geology training the astronauts endured had one primary focus: to instruct them on what sam­ples to collect and how to collect them. The training emphasized thorough verbal descriptions and proper photographic techniques to ensure good docu­mentation of the sampling site. Sampling for geological analysis on Earth has progressed to a fine art, using techniques to fit the problem under study. Proba­bly the greatest change in the past thirty years is the enormous amount of information we can now wring from a small sample (a few ounces or grams). Many of the types of analyses that let us extract this information from such small samples were in their infancy when we began planning for lunar sam­pling. But we knew that any samples brought back to Earth, no matter how small or large, would exponentially increase our knowledge of the Moon and its history. As we began to look closely at the issue and to assess the opportunities the Apollo landings would provide as well as their limitations, the sampling program became more and more sophisticated. This sophistication found its way into the types of samples wanted, the special tools needed to collect them, and the packaging or containment requirements.

Our first concern was the ‘‘grab sample’’ (later named contingency sample), one astronaut’s first order of business once he was on the lunar surface. Every­one agreed on the importance of collecting this sample in case the first EVA was curtailed, but there was little agreement on how much should be collected, how and where it would be collected, how it would be documented, what tool(s) would be used, how it would be packaged (at one point someone suggested using a spare urine bag), where it would be stowed in the LM and the command and service module (CSM), and on and on. We first thought this sample should be passed back to the astronaut in the LM to ensure that something would be returned regardless of the outcome of the landing. This operation would mean using a significant part of the first EVA time to collect the contingency sample. These concerns held not only for the first landing but for all subsequent land­ings as well. In September 1967, after a review of the preliminary timelines at MSC, I raised these issues with Mueller’s office, urging that they be addressed as soon as possible so we could proceed with tool and sample container design, which would in turn affect astronaut training and schedule development.7

Our next concern was the design of the large containers that would hold the samples on the return to Earth. They would have to be stowed in the LM on the outbound passage, then transferred to the CM for the return. Finding stowage space limited their size and weight and also their location relative to the space­craft’s center of gravity, since their weight would differ outbound and during landing maneuvers, during LM takeoff and on the CSM’s return from the Moon. Heavy aluminum boxes, called Apollo lunar sample return containers (ALSRCs), or ‘‘rock boxes,’’ were finally selected to satisfy these constraints.8 They were designed and manufactured by Union Carbide at the Atomic Energy Commission’s Y-12 plant at Oak Ridge, Tennessee. Each box weighed thirteen pounds and had an inner volume of less than one cubic foot, with outside di­mensions of approximately 19 X 11 X 8 inches. They were designed to with­stand fifty gs and to maintain a vacuum seal in case of a hard landing in the ocean. Depending on the type of samples collected, each box could hold twenty to forty pounds of material. Two boxes would be carried on each mission, and after the samples were placed inside they could be sealed while on the lunar surface. The contract with Union Carbide called for the manufacture of twelve items of flight equipment and nine test containers. Two more flight containers were added later to the contract. When the boxes were opened at the LRL, high vacuums were always found, relieving some of the worry on the first three missions that alien organisms might have escaped into the Earth’s atmosphere.

For collecting the contingency sample, a special tool was made with a long handle and attached bag. After the bag was filled, the handle would be discon­nected and the bag placed in an astronaut’s pocket in case they had to make a quick departure (thus resolving the question of spending time to get it back into the LM). With this limitation, small contingency samples were collected on each mission, always close to the LM, without much regard for the location, and not always documented with a photograph. After the contingency sample was safely in the astronaut’s pocket, subsequent sampling became much more ex­acting. Depending on the mission and the prescribed timeline, further sam­pling might be postponed until later in the first and subsequent EVAs. This later sampling would be carefully planned to ensure that the landing site was covered as completely as possible within the radius of operations.

Another concern was what type of contamination would be introduced to the samples during landing by the LM descent stage engine exhaust. The ex­haust, plus the astronauts’ activities once they exited the LM, might introduce carbon compounds, making it hard to tell if any form of life existed on the Moon. In the summer of 1965 MSC gave Grumman (the LM manufacturer) and Arthur D. Little a small contract to study these questions. In November they briefed us on what they had determined.9 There would, of course, be some contamination, estimated to be as much as one ton of various compounds spread over the landing site if they were all absorbed on the lunar surface. But chemical reactions could be predicted based on educated guesses about the composition of lunar soil, and they thought the contaminant molecules intro­duced by the exhaust could be identified during analysis of the lunar samples back on Earth. This study satisfied some, but not everyone, that the problem was understood, in particular the question of contamination from the astro­nauts’ space suits.

Concern that the samples returned might harbor some unknown disease, and the opposite fear that the astronauts might contaminate the samples on the Moon, led to the development of a sampling device called the aseptic sampler. Its function was to retrieve a small sample from an area away from the landing site, where there would be a minimum chance that the exhaust from the LM descent engine would have introduced foreign material into the soil. The asep­tic sampler was also designed and built by Union Carbide at the Y-12 plant, to specifications dictated by the National Academy of Sciences report on back- contamination. Its design became rather complicated. An extension handle would place a small coring tube against the surface a few feet from the ‘‘dirty’’ astronaut in his pressure suit. Two extendable feet would be unfolded to steady the sampler, and the astronaut would then pull a wire to open the coring device and push it into the soil. Surrounding the lower part of the handle was a sterile plastic bag into which the small core tube would be retracted; then the bag would be sealed to avoid any contamination after collection. All these functions were designed to avoid any contact with the astronauts or their gloves, because back on Earth the sample would be studied to detect organic compounds at a level of a few parts per million.

Dick Green, the ALSEP engineer and an office colleague, recalls being pres­ent at the final aseptic sampler training rehearsal by the Apollo 11 astronauts. Sam Phillips was also there to witness the demonstration of another late addi­tion to the astronauts’ workload, a sore point with NASA management (which undoubtedly prompted Phillips’s attendance). As might be expected, the com­plicated device malfunctioned. Phillips made an instant management decision to remove it from the flight and said contamination concerns would have to be resolved by studying the other returned samples (they were).

For the Apollo 12 mission and subsequent ones, two new types of samples somewhat satisfied the requirements addressed by the aseptic sample: the spe­cial environmental sample and the gas analysis sample. But there was no at­tempt to isolate these samples as carefully as if the aseptic sampler had per­formed successfully. The special environmental sample was a small container, large enough to insert a drive tube; it was taken to the Moon tightly sealed to prevent any contamination during the outbound trip. Once a drive tube sample was retrieved on the lunar surface, the container would be opened, the tube inserted, and the container carefully resealed. The gas analysis sample was designed to obtain an uncontaminated sample of any constituents of the ten­uous lunar atmosphere. The container was vacuum sealed on Earth and opened only after the astronauts were on the lunar surface. It would remain open for one or more EVAs, have a small amount of soil added, then be resealed, in hopes of capturing a few atoms or molecules that might be present in the near vacuum on the Moon.

To accommodate the procedures called for by the Field Geology Team and other scientists, several types of sample bags were designed. They would be modified as we learned from the experience of the astronauts using them on the lunar surface and the teams handling the samples back on Earth. In addition to the small Teflon bag that held the contingency sample, three other types of Teflon bags were designed to hold samples designated selected sample, docu­mented sample, and tote bag sample.

The bags for the selected sample (which replaced the bulk sample collected on Apollo 11) could contain a large volume of sample and have enough space to store the core tubes plus the lunar environment and gas analysis samples. The smaller documented sample bags (seven and a half by eight inches) were carried on a twenty-bag dispenser and would be removed individually to hold samples documented by the astronauts’ description and photographs. Each bag was premarked with an identification number that would be relayed back to MSC as the bag was filled to obviate confusion when the sample was opened at the LRL. After the selected and documented sample bags were sealed, they were placed in the ALSRCs. The large tote bag would hold any large rocks the astronauts collected. This bag would not be placed in an ALSRC but would be separately stowed, first in the LM and then in the CM.

Cameras had been part of the astronauts’ equipment since the first Mercury flights. From Gemini flight GT-4 on, they were included in formal experiments. Some good science had resulted from the pictures of Earth taken during the Gemini flights, especially new views of important terrestrial features such as the Himalayas and impact craters never before photographed.10 Cameras would become an essential element in each Apollo mission to preserve what the astro­nauts saw on the lunar surface and in lunar orbit.

On the Moon, cameras were needed for three purposes: to document the individual samples collected; to provide detailed views of the areas where the astronauts were working as well as panoramic views; and to record the place­ment of the ALSEP central station and experiments and of any other experi­ments the astronauts deployed. The Hasselblad camera, which all the astronauts were used to and which was already qualified for space flight, was an immediate candidate for lunar surface photography. Other types of cameras would be added in the months ahead, but the Hasselblad soon became the top choice.

Shoemaker and his Field Geology Team also believed that stereoscopic pho­tographs were necessary to document samples and the general geological scene.

He enlisted Homer Newell, who agreed and wrote to George Mueller that they were ‘‘a necessity on every lunar landing mission.’’11 In the summer of 1966 the Manned Space Flight Experiments Board asked Shoemaker to develop the spec­ifications for a stereo camera. Preliminary work was carried out to develop such a camera, but it was eventually canceled because of payload weight and EVA time constraints. The astronauts were then trained to use the Hasselblads to take stereo pairs.

Integrating the cameras with the astronauts’ activities became a major chal­lenge. They had to be handy but not in the way. How would the astronauts carry, point, and trigger them in their space suits and clumsy gloves? After many trials and errors, the solution was to mount the cameras on the astro­nauts’ remote control unit, a fixture attached at chest level on the outside of the pressure suit. A dovetail bracket on the remote control unit allowed the astro­nauts to slip the cameras on or off with some ease. Test subjects and the astronauts soon became adept at pointing the cameras and compensating for the parallax caused by the camera’s being below their line of sight. Camera controls were modified to be used with gloves. Once this camera was accepted, most of the simulations and training sessions included the Hasselblads, to determine how best to document the projected lunar surface activities and to get the astronauts used to them.

The camera inventory carried in the LM for use on the lunar surface was extensive. One television camera, three 70 mm Hasselblads (two with 60 mm lenses and one with a 500 mm lens), one 16 mm Mauer sequence camera mounted in the LM pilot’s window (to photograph the landing, initial surface activities at the foot of the LM ladder, and rendezvous maneuvers with the CSM in lunar orbit; it was used in later missions on the lunar surface), and about twenty-five film magazines of various types. A seventh camera, the Apollo lunar surface close-up camera (ALSCC), was one of the late additions to the science equipment.

The ALSCC was Tommy Gold’s last attempt to reap some fame from the lunar landings. Still obsessed with the nature of the fine material that con­stituted the lunar soil, he proposed a special camera to take close-up ster­eoscopic photographs of it. He submitted a proposal in 1968, and after some debate on its merits, the SSSC finally agreed to carry his camera. Shoemaker and the Field Geology Team were incensed at this decision, believing it had little scientific merit and, most important, would take time on Apollo 11 and the next missions from the much more important geological tasks and the sampling. Our office supported Shoemaker’s reasoning. We also knew that we would be assigned to oversee the rapid development of the camera while dealing with a potentially difficult PI. We were overruled, and the camera development went forward.

Gold’s photographic objectives required a complicated design for an entirely new type of camera. He wanted the camera’s focal plane to be very short, in lieu of magnifying lenses, so that particles of 0.1 mm or even smaller could be distinguished and measured; achieving this called for taking stereoscopic pairs with the camera close to the lunar surface. Since the astronauts could not bend low enough to set a camera on the surface and operate it, the camera would have to be attached to a long handle. With the camera essentially in direct contact with the surface, a light source would also have to be provided to flash for each stereo pair. On and on went the design requirements for this strange contraption that few favored, including the astronauts, who were vocal in their objections to using it. So much for the politics of science—Tommy had friends in high places.

To add to the complications, when the NASA Procurement Office learned of our plans to get bids to design and build the camera they insisted it be made a ‘‘small business’’ contract. The government’s policy of giving contracts to small businesses deserves support, and my government career after I left NASA de­pended on small business for its success, but this was a bad decision that we knew would give us trouble. Schedules were tight, and the camera’s design would require some clever engineering. We scrambled around and finally lo­cated a company (its name escapes me), and MSC awarded a contract. Robert Jones at MSC was named program manager. After several months of monitor­ing the company’s progress, it became clear that it would be unable to deliver the camera on schedule, if ever.

Now we were in real trouble, since the camera was scheduled to be carried on the first landing mission and we had lost almost six months. But because of the tight schedule, in January 1969 we were able to justify awarding a sole-source contract to the most qualified supplier, Eastman Kodak. Kodak worked literally around the clock and delivered the flight hardware and training cameras on schedule to meet the Apollo 11 launch date. Gold’s camera performed almost flawlessly, thanks to the Kodak engineers, and it was also carried on Apollo 12 and Apollo 14. Although it was not a favorite experiment for the astronauts—a few threatened to throw the camera away—they complied with most of his requests for his unusual photographic subjects and returned forty-nine and a half stereo pairs.

How much new science resulted from analysis of the photographs is debat­able. Gold tried to use them to advance some of his pet theories, and David Carrier, an MSC engineer who had provided oversight on the soil mechanics experiment, reminded me that when he and several other MSC staffers cooper­ated with Gold in writing his report for Apollo 14 they withdrew their names as coauthors because they disagreed with some of his conclusions.

When more weight became available on the J missions, the tool inventory remained essentially the same except that we added a rake, suggested by Lee Silver after the Apollo 12 mission when the astronauts found it difficult to pick up small rocks and collect samples mixed with the lunar soil. We reasoned that such samples would yield a wide variety of lunar rocks, since every landing site might contain ejecta from many distant sources. The rake was designed as a scoop, closed at one end, with wire tines spaced about a quarter inch apart to sift out the loose material but retain the larger pieces. It was used successfully on all three J missions.

We added another important piece of equipment for the J missions, the Apollo lunar surface drill. Two requirements led to its development: the ALSEP heat flow experiment, which needed two holes for inserting the sensors, and the geologists’ and geophysicists’ desire to obtain subsurface samples. Here once again the experience gained in studying a deep drill for the post-Apollo mis­sions was valuable. Jack Hanley, detailed to my office from USGS, had moni­tored the hundred-foot-drill studies at Marshall Space Flight Center, and he was assigned to oversee the drill. The RFP released by MSC called for bids to build a drill that would extract cores to a depth of one hundred inches. The competi­tion was won by Martin Marietta, Denver, teamed with Black and Decker.

The design the Martin Marietta team selected was a battery-powered rotary percussive drill in which the power head imparted short impacts at the same time as the drill pipe (core stem) rotated. The astronaut could also lean on the drill handle to add force and improve the penetration rate. The core stems (a total of six that would be screwed together during the drilling) were fluted on the outside, as in the hundred-foot drill studied by Westinghouse several years earlier, to carry the cuttings or soil to the surface as the drill penetrated into the subsurface. Each core stem, made of fiberglass tubular sections reinforced with boron filaments, was about sixteen inches long. As each one penetrated to its full length, the drill head would be disconnected and another core stem screwed on to continue drilling. A tripod device held the extra sections above the ground until they were connected during the drilling. There was enough bat­tery power to drill three holes: two for the heat flow experiment and one for the core sample.

After five Surveyor spacecraft had landed on the Moon and returned pic­tures and rudimentary data on the characteristics of the lunar surface, many questions still remained about some of the engineering properties of the upper layers of the lunar surface. Since the Surveyor spacecraft had not disappeared in fluffy dust, we now knew that traveling on the lunar surface in some sort of wheeled vehicle would be possible. Using lunar soil to shield shelters while lunar bases were being built (as proposed in the Lunar Exploration System for Apollo studies) also appeared feasible, but more hard data were needed to understand how these soils could be excavated.

The need to predict the behavior of lunar soil, insofar as it would affect the design of vehicles and other equipment, as well as the need to collect other basic information, led to the inclusion of a soil mechanics investigation on the final four Apollo missions. This experiment, closely allied to the field geology stud­ies, consisted of analyzing the astronauts’ observations on the character of the soil as they moved about; photographing the soil after it was disturbed by their activities (e. g., boot prints, tire marks, and trenches), augmented by physical measurements made in situ with penetrometers and other devices; and finally, making measurements on the returned samples.

James Mitchell, from the University of California, Berkeley, was selected as the soil mechanics principal investigator. His team included as coinvestigators Nicholas Costes from MSFC, who had been on the Apollo 11 and Apollo 12 Field Geology Team and had participated in some of our post-Apollo studies, and Dave Carrier from MSC. Don Senich, a former instructor at the Colorado School of Mines who was detailed to my office from the United States Army Corps of Engineers, was to oversee the development of this experiment from headquarters.

A simple penetrometer, consisting of a long aluminum shaft slightly less than half an inch in diameter, was carried for the first time on Apollo 14. It was to be pushed into the surface at several places near the LM to a maximum depth of sixty-eight centimeters. Black and white stripes were painted on the shaft,

and after pushing it as deep as possible each time, the astronaut would read back the number of stripes still above the surface as a measure of the depth achieved. Mitchell’s team would then calculate the forces involved by applying data obtained from terrestrial simulations. On the Apollo 15 and Apollo 16 missions a more sophisticated, self-recording penetrometer was carried. This device consisted of a base plate, a shaft with two different-sized interchangeable nose cones, and an upper housing containing the recorder. An extension handle above the recorder helped the astronauts force the nose cones into the surface. After pushing the penetrometer into the soil, they would remove the data drum from the recorder and return it for analysis.

Chapters 11 and 12 will tell more about how the equipment for the field geology experiment was used on the Moon by the crews of the six landing missions.