SSI Newsletters: 1988 March April

Space Studies Institute Newsletter 1988 March April cover

SPACE STUDIES INSTITUTE
P.O. BOX 82
PRINCETON, NEW JERSEY 08540
[[librarian note:  This address is here, as it was in the original printed newsletter, for historical reasons.  It is no longer the physical address of SSI. For contributions, please see this page]]

 

SSI UPDATE
THE HIGH FRONTIER® NEWSLETTER
VOLUME XIIV ISSUE 2
MARCH/APRIL 1988

 

VICE PRESIDENT’S COLUMN

This month’s column is being written at the Atlanta airport, where I am waiting to catch a plane to the Marshall Spaceflight Center in Huntsville, Alabama. There, Major Alex Gimarc, Ron Jones and I will be briefed by Martin Marietta Corporation and NASA on the possible conversion of a Shuttle External Tank to a gamma ray imaging telescope. Of particular interest to the Institute is the fact that over the past year NASA and Martin have conducted experiments on all sorts of challenges that astronauts will face in using and modifying the tanks on orbit. The preliminary word is that tests using the neutral buoyancy chamber revealed no show-stoppers.

New National Space Policy

As we go to press, the President is unveiling a new national space policy, which appears to be based in large part upon the recommendations of the National Commission on Space. According to early releases, one element of this policy is that commercial entities will be able to obtain Shuttle External Tanks. In addition, there was a strong emphasis on supporting commercial space activities. Indications are that this new policy will point toward future lunar outposts and the technologies required to create them. Watch the SSI Bulletin Board, (xxx-xxx-xxx), and future issues of “Update” for the latest developments.

Lunar Systems Study

The President’s new space policy will come as welcome news to participants in SSI’s Lunar Systems Workshop conducted in January. Our team of over 20 researchers met to examine what activities in space can provide the commercial payback needed to bootstrap the large scale use of space resources and fuel the human breakout into space. The results of the study are now being documented and will be presented at the “Lunar Bases” conference in Houston, (see page 6). The study results will be the subject of a future article in “Update.”

Other SSI Projects

The space resources computer modelling team reports that they have analyzed the surveys taken of SSI’s Senior Associates and at our Princeton Conference and are in the final process of selecting the language in which the model will be authored.
Dr. Les Snively now has a mass-driver simulation operational on the Macintosh. He is comparing results of simulated runs with actual data from Mass-Driver III hardware.

SSI has met with a number of space organizations with regard to the lunar polar prospector. NASA/JPL is exploring the use of a surplus Atlas-E to do a quick low cost lunar polar mission. Dr. Gerard O’Neill and I briefed NASA Headquarters on this and other low cost lunar polar probe options last fall.

Brandt Goldsworthy and his team have completed the construction of a special press which will be used to create the first glass/glass composites from simulated lunar material. Goldsworthy’s paper on SSI’s composites research has been accepted for presentation at the “Lunar Bases” Conference in April.

Ronald Jones has completed an excellent slide depiction of an External Tank-based variable gravity facility. Ron Jones’ External Tank slides were shown at a recent briefing at the Marshall Spaceflight Center and were given a huge ovation by Center personnel and visiting dignitaries. Bill Engfer reports that the tanker toys AUTOCAD representation of the Shuttle External Tanks is nearing completion. These computer files will enable AUTOCAD users to explore External Tank designs. For further information contact SSI Headquarters.
The Institute has established an all­ volunteer Teleoperations Working Group. (See article on Page 5.)

Space Studies Institute Newsletter 1988 MarFeb image 4
Space Studies Institute Newsletter 1988 MarApr image 5
Space Studies Institute Newsletter 1988 MarApr image 6

We have selected a volunteer intern from Princeton University to conduct experiments in our lunar mining simulator. Audrey Robinson, a mechanical engineering student, will spend a portion of March working in our new facility.

Gregg Maryniak

 

GUEST COLUMN

THE ISSUE IS COST
W. Haynes, Senior Analyst
Science Applications International Corp.

Statement of the Problem

We are concerned with the cost of launch, but the cost of all government-funded high technology, especially space-related technology, is much too high. The cost levels generated by the government create an environment and an industry that can not reduce costs for equivalent private sector customers because they only know one way to operate and that is the governmental way. In addition, it actually becomes unlawful to use lower overhead rates for a civil (commercial) customer than for a government buyer in the same plant or cost center. The reasoning is if that were allowed, added overhead would be billed to the government customer in order to be more competitive in the commercial market.

Why is government procurement so expensive?

There are many detailed reasons, such as safeguards to assure safety, to control costs, and to expedite contract completion. These take the form of specifications and clauses in procurement regulations cited in the front of contracts and administered by the government procuring authorities.

The greatest irony is that recent events have demonstrated that our systems are neither safe, economical nor timely. They have been failing at an unprecedented rate, are exorbitantly costly and are taking ever longer to acquire.

Remedying these problems in a piecemeal manner by searching out individual requirements and eliminating or reducing their impact can not be successful, because each regulation and requirement will have its staunch advocates who will fight to retain it, using well founded arguments such as safety, economy and the prevention of malfeasance and misappropriation. Never mind that the aggregate of requirements has failed to achieve these noble goals in the past. It will prove impossible to prove that the particular regulation or directive was faulty in order to eliminate it. An example exists in the form of the recommendations of the Rogers Commission after the Challenger accident investigation. They proposed that NASA add more safety, maintainability, reliability and quality assurance people and procedures. NASA has added an Associate Administrator for R,S,M and QA and 200 additional safety inspectors. This is in spite of the fact that the deficiency which caused the accident was not unknown to NASA management, but had been clearly cited by the responsible engineers months before and on the night before the launch as a potential catastrophic failure mode.

Therefore a method for reducing costs that does not involve detailed procedural change must be used.

Background

In July, 1986 a paper was given at the Interagency Conference on Environmental Systems in San Diego. It described a “man in a can” system for allowing an astronaut to move around in space autonomously. The “can” was to have reaction control rockets, two manipulators operated by the astronaut inside the can and a fixture to allow the operator to stabilize the system by docking it to available structures. The interior provided a shirt-sleeve atmosphere at sea level pressure.

The estimated budget through first article delivery was $410 million.

No one blanched when that figure was announced.

Space Studies Institute Newsletter 1988 MarApr image 1
The presenter had provided a picture of a device called “Deep Rover” as an example of an equivalent system designed for use underwater. I was acquainted with Deep Rover and with its designer, Phil Nuytten, president and owner of Canadian Divers, Inc. (CanDive) of Vancouver. Deep Rover consists of a six foot acrylic sphere, about six inches thick and equipped with remote manipulators, six degree of freedom maneuverability, navigation and control systems allowing full, untethered mobility between the surface of the ocean and depths down to 3,000 ft. I wrote to Phil Nuytten and received confirmation of what recalled from personal conversations: He designed, fabricated, tested and delivered Deep Rover for $1.8 million dollars! Furthermore, he said he can deliver additional units for $1 million each and make money on them, but he prefers to lease them out to the petroleum industry with operators. The unit, as built, includes all servicing equipment, hoists, shelters for housing the system on the fantail of the using vessel and installed safety equipment for the one man crew. He told me the vehicle is steered by a small microprocessor using a simple joystick. Checking out an operator takes about twenty minutes, he said.

The corresponding space system would operate with only a one atmosphere pressure differential, with the low pressure on the outside. The manipulators, RCS, life support systems, communication systems, and attitude control electronics for the space system should be available from existing systems such as the Manned Maneuvering Unit, JPL and Ames Research Center prototype arms and the Orbital Maneuvering Unit being designed by TRW. Yet it will cost over two hundred times as much as the Deep Rover, according to the designer and would-be manufacturer.

Deep Rover must operate with over 100 atmospheres of pressure differential, with the high pressure on the outside, and contend with currents, corrosion, rough handling on a pit­ching deck, and a generally more hostile environment than would the “can.”

Why is the projected cost differential so high in favor of Deep Rover?

The answer is not in the technology ­ both technologies are mature and off the shelf, nor in the environment (Deep Rover’s is much more difficult). It must lie in the design and fabrication (the acquisition) process.

Nuytten had a small design team. They produced the minimum in drawings and system definitions, procured from well established sources on a relatively informal basis and hand built the system from scratch in their own facility, using a lot of finished parts. The first article was also the operational article. Instead of a NASTRAN Analysis to verify the structural integrity, for example, they depended on the known qualities of the acrylic and built in adequate safety margins. They then got in it and checked it out in the water, gradually increasing depth and maneuvering range.

There is nothing about that process that is unique or particularly innovative, nor is there anything that would preclude using a similar approach to building the “man in a can” system or many other space systems except that “we have never done it that way.”

Discussion

The difference lies in the procurement process and in the design, fabrication and testing processes that the procurement process dictates. There is some evidence that the procurement process also affects the fundamental design approach itself. That seems to result in technical overkill in some instances and in suppression of innovation in others.

Nuytten knew what he wanted in the way of performance and he set out to build a system that would provide him with it. The only specifications he had to adhere to were those dictated by common sense, the market, and the economics of the situation, and the fact that he was personally going to travel 3,000 ft under water in the finished product. Surely that fact operated as an adequate incentive to build in the requisite reliability and safety; far more than application of abstract and generic safety specifications imposed by routine contractual fiat.

The application of a similar approach to procurement of high tech systems for the government seems to offer opportunities for radical cost reductions. What would happen if the government announced that it required a device for transporting an astronaut between orbital facilities in a “shirt sleeve” environment, with the device to be provided for a period of ten years, including delivery to a specified orbit, guaranteed availability for specified hours per year and asked for yearly leasing bids? I predict that the mainline aerospace firms would proceed as they are wont to and produce a system that would cost in the hundreds of millions of dollars. Given the proper encouragement and assurances that the government really meant what it was saying, firms such as CanDive would be able to meet the need with systems costing a very small fraction of what would be required by the conventional firms. Those systems would be tested in space by the firm’s own employees and maintained by them. The rental would cover the firm’s costs in the same manner as any terrestrial leasing firm, and could represent a very substantial reduction in costs to the government.

The practical problem would be preventing the government representatives from imposing the same environment upon the lessor that has been imposed on all of the aerospace manufacturers (with their aid and support) to date. The professional astronauts would protest that the commercial space workers were not adequately equipped to evaluate the requirements for designing, building, or operating such a system, nor to operate it. They would insist on validating the system and produce a long list of “improvements” that they would consider essential to assure safety and reliability. They would point out NASA regulations requiring specific testing and qualifications for any system that is to fly in space, and claim that the lack of imposition of those procedures could jeopardize the Space Shuttle which must deliver the system to orbit. The ultimate result would be that costs would be driven right back up to almost the level they reach under normal government procurement and we would be back where we started, except that the innovative civil procurement process would have been discredited with no solution to the problem of excessive costs.

This points out the necessity for establishing a complete civil, commercial space architecture if the benefits of commercial standards and procedures are to be realized.

Another example of a system that was inicitiated as a commercial effort to reduce costs is the Spacehab mid-deck extension module. The Spacehab module is the brainchild of Robert Citron of Seattle. He founded a private company to design, build and market this device which will be carried in the forward volume of the Orbiter cargo bay. It will provide about 1000 cu ft of additional pressurized volume to enable many more shirtsleeve experiments to be conducted. Spacehab, Inc. has received enthusiastic response from the international industrial and scientific community with many commitments to lease space when the system flies. Unfortunately, the system is totally dependent upon the Space Shuttle for flight, and NASA recently reneged on a previous statement of intent and postponed first flight well into the 90’s. That was in spite of many NASA scientific experiments and researchers who wished to book space on the Spacehab. Fortunately, Congressional pressure has resulted in a reversal of that stifling decision, and NASA has now agreed that the Spacehab module will be launched on a relatively near term Shuttle flight so that it may be used to reduce the backlog of scientific experiments created by the Challenger accident and the subsequent launch moratorium.

Similarly, I was recently informed that NASA was withdrawing its commitment toe launch the Industrial Space Facility being designed and built as a private enterprise mini space station by Space Industries, Inc. That decision was also reversed by Congressional pressure.

These actions are a clear measure of NASA’s real attitude towards “competitive” commercial projects and bode ill for efforts to procure directly from commercial operators as long as NASA has the ability to choke off access to space.

The Booster Problem

An article in Newsweek by Gregg Easterbrook entitled “Big, Dumb Rockets” has aroused a great deal of discussion. He referred to the efforts of an Aerospace Corporation engineer to persuade DOD and NASA managers to develop simpler systems and avoid the excessive costs of the hi tech systems they seem to favor. He used the Space Shuttle as an example.

Although many of the points made in the article are, in my opinion, valid, the writer is carried away by his own rhetoric and slips into an acrimonious style. He is also internally inconsistent. Easterbrook produced a number of cost figures without adequately defining their basis. In referring to Shuttle costs, he cites figures which were originally quoted for a Shuttle with a fly-back first stage (a hi tech solution with relatively high up-front costs) and then shows how the current Shuttle costs vastly more. He fails to state that the current (lower tech, only partially re-usable Shuttle) was forced on NASA by the arbitrary halving of the Shuttle budget by the Nixon administration. In other words, the Nixon administration employed exactly the approach Easterbrook advocates in his article (they forced NASA to “economize” with “dumb boosters” for the Shuttle instead of developing the hi tech fly­back first stage) and the result has been very substantial increases in operating (recurring) costs. That has locked the Shuttle into a situation where the system cannot reduce life cycle costs because of the high recurring costs. Had NASA been allowed to spend the up-front money to develop the fly-back first stage, in­tegration and processing costs would have been substantially reduced. The recovery fleet for the solid rocket motors, the reprocessing costs and the manpower needed for re-integrating the expendable External Tank, solid motors and the Orbiter, would then all be eliminated and a relatively straightforward mating of two flying vehicles substituted. The life cycle costs for payload delivery to orbit would be decreasing steadily as the number of flights and the payload mass delivered increased. As it is, the solid (dumb) boosters not only cost more to integrate, they are also the cause of the Challenger tragedy. That creates a credible argument that it was the Nixon administration’s decision and NASA’s failure to stand up and resist that edict that led directly to the Challenger accident. The connection is not, however as direct as the facts stated above seem to indicate. My perception is that the sequence of cause and effect is more subtle. Once the budget reduction was imposed and NASA began to adapt to the resulting austerity, a pattern of deferral of problems became established. For example, the Orbiters never had an adequate allocation of main engines, necessitating cannibalization between vehicles and excessive engine change-outs.

Shortly after STS-1, I was speaking with Dean Grimm, a senior NASA JSC engineering manager who has since retired. He told me that after STS-1 landed, NASA called the manufacturer of the turbopumps used on the space shuttle main engines. They told the manufacturer that two of the six turbopumps had recorded vibration anomalies just before main engine cut off, and that they wanted the manufacturer to conduct a teardown inspection of those pumps, prepare a report on their condition and restore them to a zero-time condition. They also wanted to order some spare pumps. He said there was a long pause on the other end of the line, and then the manufacturer’s representative said: “Apparently you’ve forgotten, but you terminated our contract for those pumps as an economy measure last year. The pumps were manufactured at a small plant in Alabama which we had to shut down. The employees are retired or transferred, the tooling has been scrapped and we don’t even have any drawings of those pumps. Now, would you like us to start all over again?”

When NASA made the decision to terminate that contract they were not being stupid. Dean Grimm told me that they were faced with a very simple choice: they had sufficient budget to finish the Shuttle and fly it, or to provide adequate spares, but not to do both. When the NASA managers were faced with a potential but not yet clearly imminent problem, such as eroding O-rings, long practice resulted in the problem being deferred along with a host of similar ones. The similar ones included marginal APUs (auxiliary power units), marginal wheel brakes, antiquated on-board computers, the aforementioned main turbopumps, a single-string reaction control system and others. That is a reason for the extended grounding resulting from the Challenger accident: NASA finally has been able to face up to the approximately 2000 failure mode effects analyses/critical items list that had accumulated before the accident.

The fact is, the 0-rings just happened to be the first thing to fail catastrophically. That is more of a tribute to the redundancy NASA had built into the space transportation system than to the reliability of the subsystems. The rule for all systems has been that they must be “fail operational, fail operational, fail safe” if their failure would be life-threatening, or a deviation would have to be requested and granted. An example is the APU system. That system is essential to the operation of the Orbiter because it supplies the hydraulic power to operate the aerodynamic flight controls used during atmospheric flight and in the approach and landing. There are three APUs on board, and on at least one flight, one APU was out and a second was “limping” when the crew was directed to return and land prematurely. In other words, the APU system is fo/fs, and requires a deviation or a fourth APU. The basic structure of the vehicle must, in many cases, be exempted from the fo/fo/fs requirement because it simply is not possible to meet that requirement.

Admittedly, many of these problems are a product of the STS design. It is generally agreed that, given the chance, NASA would design an all-electric Shuttle and eliminate the APUs in favor of additional fuel cell electrical power. The AP-101 IBM computers are being replaced. Imagine our leading spacecraft being controlled by 1968 vintage computers? Remember circuit boards with discrete resistors and capacitors and programming with binary machme language?. The major expense is not, however, the computers: it is the software. The Shuttle software has been developed, flight proven over the first twenty five flights, and there is great confidence in it. Rewriting and then flight-proving it is not a chore to be taken lightly, or cheaply. It also requires a change in all of the supporting ground equipment including simulators, the shuttle avionics integration laboratory and interfacing command and control hard and software. One of the favorable results of the Challenger accident is that many (but not all) of these problems can now be dealt with.

Having said all of this, the fact is that launching payloads on the STS is indeed far too expensive, whether one accepts Easterbrook’s figure of $6,000/lb as actual amortized life-cycle cost or the NASA figure of $2,800. The latter figure was arrived at by dividing per flight costs of $180 million by the design payload weight of 65,000 lbs. Current published lift capability post-Challenger is 52,000 lbs, resulting in a $3,400/lb figure. The OMB has tried in vain to determine actual fully amortized, life cycle Shuttle per lb costs. The problem becomes an argument over accounting practices and how to allocate pro-rata shares of NASA overhead to the STS.

The U.S.A.F. at the recent advanced launch system bidders conference, announced that the ALS would have to achieve a life cycle cost per pound delivered to orbit that was an order of magnitude lower than current systems. Then the Air Staff people who conducted the briefing stated that the Titan IV, which will fly for the first time in 1988, is the bench mark and that its life cycle costs have been estimated as $3,600/lb. The ALS must achieve $360 per lb. That was in fact, the only hard requirement stated by the USAF. They were explicit that the life cycle costs had to include everything associated with design, fabrication, test, delivery and operation of ALS, including tooling, brick and mortar, etc. They did not define any hardware approach, not even whether it should be expendable, re­usable, or a hybrid, manned or unmanned ­ nor did they specify a payload capacity per flight. In fact, the ALS study contracts are a strong move away from the detailed specification of system requirements approach which has been standard for DOD as well as NASA.

Unfortunately, the ultimate goal appears to be to have the contractors define a system, after which the DOD, NASA and SDIO (the three co-sponsors of the studies) will choose a design and proceed as usual to purchase the hardware rather than the service. Of course, if ALS actually achieves the order of magnitude reduction in costs specified, or even comes close, it will rapidly replace all other launch systems worldwide. This is especially so if there is a manned option and the ALS can do those jobs which, up to now, only the Shuttle could do, such as deliver man and return him from orbit, and return other payload from orbit.

I have examined the implications of the cost goal of ALS and arrived at some preliminary conclusions.

At $360/lb life cycle costs, if we assume an even split between non-recurring and recurring costs, 100,000 lbs payload rating and 24 launches per year for ten years, the allowable total budget for recurring costs will be:

$360/2 * 100,000 * 24 * 10 = $4.3 billion

More realistically, if we assume an 80% average mass load factor (the STS has averaged less than a 60% load factor based on the nominal 65,000 lb payload) half of the budget drops to $3.45 billion. For 240 launches, that comes to about $14.4 million per launch. That seems unrealistic for a conventional rocket system. Assuming a 25/75 per­cent split would allow $1.7 billion for non-recurring costs, $5.2 billion for operations (recurring costs) and a budget of $21.6 million per launch for 240 launches over ten years.

Clearly, it will be difficult if not impossible to achieve that figure with an expendable system. Even a fully reusable system will have to be very innovative to meet that goal.

The goal becomes more reasonable as the cost split is heavier towards the non-recurring, provided that the resulting operations budget is reasonable in the first place. That is because the marginal cost of launching additional payload is less than the fixed (up front) costs. That is a lesson we should have learned and applied to the STS; if we had, we could at least have argued for the (hi tech) flyback first stage, although government is rarely interested in saving money in the long run at the price of increasing current budgets.

At this point someone is bound to point out that another way to reduce life cycle costs per pound is to increase the payload capacity per launch because the marginal cost of building a bigger rocket is less than the cost of another launch with a smaller rocket. That reasoning, however, is incomplete.

The 200,000 lb lift capacity rocket is indeed more efficient than an identical smaller counterpart as long as you have adequate demand for launching 200,000 lb payloads. That is born out by the cost per lb of the small Scout rocket, which is reported as $12,000.

The problems begin when you run out of single 200,000 lb objects to orbit. Now you must aggregate 5, 10, and 15,000 lb payloads to total the rated payload capacity of the heavy lift launcher. That creates a host of problems. The aggregation will require a hardback to support each payload, as they can not simply be stacked on each other. That hardback will weigh on the order of 15 or 20,000 lbs and be peculiar to each aggregation of payloads. It will have to be designed and tested, qualified for the specific payload aggregation and will be thrown away as part of the expendable system each time. To gain an appreciation for the significance of this, recalls that one of the costly effects of the Challenger accident was that the TDRSS pallet used to stow that com sat in the Orbiter cargo bay was lost. That pallet was normally returned and reused. The hardback will be relatively intolerant of major changes in payloads and will add to the queuing problems produced by trying to assemble all those diverse payloads. The initial payload managers will grow very impatient while waiting for the aggregation to be assembled (it could take years) and will probably opt for a more expensive but available small launcher. Every time a payload defects, the hardback design will have to be altered ­ even if they decide to leave the slot vacant ­ because of the altered dynamics. Furthermore, the aggregation will hardly wish to be delivered to the same orbital parameters, necessitating use of an orbital tug to arrive at their respective final locations.

That may create increased demand for the full capacity of the heavy lift launcher in order to supply the propellants for that tug, but that merely adds inefficiency to the system as a whole.

The inevitable result will be that the super-rocket will be launched at partial capacity – with resulting erosion of the purported cost advantage. But the argument will be that we have built this monster, so now we have to use it.

A New Perspective

By applying a lesson learned by our terrestrial transportation industry, we can apply a new perspective to our space transport systems.

Surface transportation of freight and cargo is divided into two basic categories: bulk transport of handling-insensitive materials that conform to any container, and individual handling of delicate items requiring TLC. The former are represented by oil tankers, ore trains and pipelines, while the latter are represented by Federal Express and the Postal System. The former are also characterized by their delivery of cargoes to central depots for further processing and final handling, while the latter are generally sent to specific final destinations. The difference in costs is reflective of the nature of the processes, but we accept that difference because we understand the nature of the services being rendered.

To illustrate the difference in perspective, we have been operating in space as though we were using Federal Express to deliver coal! Examination of the Space Transportation Architecture Study Mission Model shows a high percentage of bulk cargo. Yet there has been no effort to separate the space cargo delivery systems as the nature of the task has separated our terrestrial delivery systems.
If we recognize this difference, we will search for the best way to meet each need, and that search may result in a more innovative approach and the potential for very significant life cycle launch cost reduction.

Proposed Studies

At a meeting of the U.S.A.F. Space Division Cost Reduction Workshop in August, 1986, about twenty of our leading rocket engineers had spent most of the day discussing how to reduce integration and launch operations costs for current systems. The conclusion seemed clear that at best, we might be able to shave 20% off current experienced costs, and that it would require several years to implement the proposed changes. I asked why we were spending so much time in this relatively unpromising effort when the goal was to achieve an order of magnitude cost reduction?

One of the discussion leaders informed me that they had about given up on achieving the factor of ten cost reduction, and that, besides, anything we came up with would be applicable to the follow-on systems, also.

I replied that the second statement explained the first; if the follow-on systems were so similar to the current systems that our limited remedies would still apply, then it was no wonder that they had given up on achieving a ten fold reduction in costs.

The point is that cost reductions will occur. Every transportation system ever conceived by man has had a history of initial high cost and eventual reduction to the point where it became accessible to the average person. There seems to be no reason why space transport should be the sole exception. What is required is to apply innovative thinking to the problem. I was asked what I had in mind. I replied that there are two possibilities which I would cite as promising: vertical take-off and landing, single stage to orbit rockets and electromagnetic space launch systems.

Since then we at SAIC have pursued in­house studies of both systems in preparation for approaching the government for funding a vigorous evaluation of their possibilities.

Both, for different reasons, exhibit the essential requirement for a system with reduced life cycle costs. Their development costs may be high, but they offer the promise of not one, but two orders of magnitude reductions in life cycle costs per lb delivered to orbit. They accomplish this by their potential for sharply reduced recurring costs, so that the increase in payload delivered rapidly amortizes the initial costs.

VTOL/SSTO

This concept has suffered from the initial “proofs” that rockets could never reach space because of the limited specific impulse of black powder. It was the invention of the multi-stage rocket which overcame that barrier. Ever since, it has been gospel among rocket engineers that SSTO is “impossible”. There have been advocates of SSTO over the years, but no serious attempts to build such a system. Nevertheless, the trends of technology are such that the SSTO is, or will become, possible before the general community of rocket engineers is ready to accept it.

In the early years of the space program President Eisenhower launched an Atlas with a radio on board which broadcast a Christmas message to the world. It was known as “Project Score”. In that case the entire Atlas rocket was inserted into orbit. Although the Atlas does separate a half-stage (two small boosters attached to its sides) this was very close to a demonstration of SSTO, and it occurred over twenty years ago.

A powerful boost was given to the SSTO cause recently when Ivan Bekey, NASA Director of Advanced Programs, suggested SSTO for the Shuttle II studies (Aviation Week & Space Technology, Dec. 1, 1986, pg 30). He also provided a list of the technical developments which he believes make SSTO feasible. The remarkable feature of Bekey’s proposal is that he proposes a horizontal landing system. That is remarkable because of the mass fraction penalty exacted by having to carry wings into orbit. If SSTO is feasible under that handicap, how much easier it must be if the vehicle lands vertically and thereby) dispenses with wings and the structure to support them.

The study that SAIC wishes to conduct proceeds on the premise that VTOL/SSTO may not be possible now, but that it will be very useful to determine by what margin it fails. That is, if a parametric analysis is conducted, by how much would currently achievable mass fractions have to improve in order to make the system viable? Would, as is claimed by some studies, the currently possible systems produce a negative payload capability? What would be necessary in order to deliver 20,000 lbs to low earth orbit with such a system?
The pay-off for VTOL/SSTO can best be appreciated by considering that a 450,000 lb gross lift off weight system would use about $200,000 worth of propellants (liquid oxygen and liquid hydrogen). If we recall that there would be no integration costs and only routine maintenance required because the vehicle is launched as a unit, and returns as a unit, we begin to see a rocket system that may truly approach airline type operational costs. If turn­around and amortization cost an additional $200,000, the cost would be $20/lb to deliver the 20,000 lbs, for direct operating costs. Cutting the cargo capacity in half would still result in only $40/lb costs. Clearly, that system will very rapidly amortize even very substantial non-recurring (development) costs. Such a system would also enjoy the ability to turn around between flights very rapidly, and to truly provide the kind of “launch on demand” called for by the DOD. It would seem worth­while to determine just how close we are to such a system now, and how soon it may become possible if we really work at it.

The VTOL/SSTO is my candidate for the TLC half of the launch fleet.

Electromagnetic Space Launch

EMSL has been advanced by vigorous research and testing of railgun systems for SDIO. A prototype battery driven railgun is being constructed under SAIC direction in Shalimar Florida, for the SDIO.

Electromagnetic space launch has been advanced by vigorous research and testing of railgun systems for the strategic defense organization. A prototype battery-driven railgun is being constructed under SAIC direction in Shalimar, Florida. Railguns have exhibited very high accelerations; however, they are low duty-cycle machines. The other major type of electromagnetic launcher is the Mass­Driver. Mass-Drivers have been constructed at M.I.T. and Princeton University under the direction of Dr. Gerard K. O’Neill and with the sponsorship of the Space Studies Institute. Mass-Drivers have been improved with accelerations ranging from 33 gravities for the first machine, built at M.I.T. by Professors O’Neill and Kolm, to 1800 gravities for Mass­Driver III constructed by the Institute at Princeton.

Another type of coaxial electromagnetic launcher, the coilgun, is being continued by Henry Kolm through his company EML, a subsidiary of Kaman Corporation.

The application of EMSL to space would fulfill the other half of the space analog to terrestrial transport: the bulk transport of conformal, acceleration-insensitive materials.

SAIC has conducted some fairly detailed preliminary analysis of the driving factors governing EMSL and is preparing a formal briefing to be given to U.S.A.F. Space Division and SDIO people in the near future.

The EMSL is envisioned as an evacuated tube through which the cargo carrier is accelerated to about 11 km/sec. It will be inclined about twenty degrees above the horizontal and be about 600 meters long, resulting in an average acceleration of about 9000 g’s. The launch tube will be buried and may have to be constructed in a remote location because of noise levels. Diameter is determined by atmospheric drag considerations and acceleration loads, two areas requiring further investigation. The g’s are not considered especially problematic because such accelerations are sustained by, for example, the Copperhead smart artillery shells which contain laser tracking and active guidance systems.

The EMSL carrier vehicle will require a circularization rocket in order to achieve a stable orbit, although the launcher exit velocity of 11 km/sec is expected to be reduced by only about 3 km/sec in the two seconds required for it to leave the atmosphere.

Many questions remain, but the essential technology appears to be emerging. The pro­posed study will evaluate feasibility of an orbital velocity and altitude demonstration launch, possibly using an upgrade of the system under construction in Florida, and preliminary definition of an IOC (initial operational capability) EMSL.

Conclusions

Achieving substantial cost reductions for space systems by pursuing low tech systems per se will not be successful. On the contrary, only by applying new technology just coming into the realm of the possible, can really substantial cost reductions be achieved.

Nevertheless, aerospace systems are much too expensive, as can be shown by comparing similar civil and governmental systems developments. The solution is in eliminating the encrustation of supposed “safeguards” which have caused the aerospace industry to evolve as it has. The best way to achieve that is to initiate a process of procuring services, as distinguished from hardware, wherever that is possible. The government must then resist the inclination to second guess the commercial operator and reimpose the traditional governmental “security blanket” lest the costs be reimposed in the process.

A vigorous and truly daring program of development of “fringe” ideas now coming into the realm of the possible offers the bast chance for real breakthroughs in both cost and performance (e.g.: pursue the USAF Project Forecast II objectives).

Finally, it is not hi tech which is the problem, it is the cost of high tech. And it is not the technologists who cause the high costs (although they aid and abet it), it is the government procurement system. The solution is to go to purchase of services, by competitive bid, FOB the specified location. The government customer should specify what services are required and then get out of the way!

 

LUNAR TELEOPERATIONS STUDY GROUP FORMED

The Space Studies Institute (SSI) has recently formed a Lunar Teleoperations Study Group, chaired by Robert B. Lewis, of General Electric’s Astro Space Division. The study group will examine the potential of teleoperations in conjunction with the Institute’s broader goal of hastening the human breakout into space.

Teleoperations is a means of providing remote control and observation in distant or hostile environments. From a human operator’s point of view, teleoperations is ‘the next best thing to being there.’ From a ‘shirt sleeve’ office environment an operator can control sophisticated robots and other specialized machines located in space or another planetary environment.

Space Studies Institute Newsletter 1988 MarApr image 2

Space Studies Institute Newsletter 1988 MarApr image 3

The study group is examining relevant research already performed and identify critical path experiments which can be used to evaluate teleoperations potential in a lunar Lunar Systems Workshop will serve as a framework for these evaluations.

Specific low cost proof of concept experiments will be performed this year to begin to verify critical technical issues. The group is open to serious suggestions for experiments and offers of assistance. Contact may be established through the SSI office. A special conference area has been established on the Institute’s computer bulletin board to post the group’s activities and foster communication with the SSI membership.

Space Studies Institute Newsletter 1988 MarApr image 7
©space studies institute


Back to the full list

NEXT: 1988 May-June (GKO’N Forewards The Overview Effect, SSI Lunar Systems study)

Space Studies Institute
7429 Laurel Canyon Blvd
Suite S
North Hollywood CA 91605
(661) 750-2774
admin@ssi.org