National Academies Press: OpenBook

Opportunities in Chemistry (1985)

Chapter: VII. Resources for Basic Research in the Chemical Sciences

« Previous: VI. Manpower and Education
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 288
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 289
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 290
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 291
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 292
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 293
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 294
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 295
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 296
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 297
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 298
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 299
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 300
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 301
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 302
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 303
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 304
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 305
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 306
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 307
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 308
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 309
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 310
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 311
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 312
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 313
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 314
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 315
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 316
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 317
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 318
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 319
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 320
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 321
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 322
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 323
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 324
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 325
Suggested Citation:"VII. Resources for Basic Research in the Chemical Sciences." National Research Council. 1985. Opportunities in Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/606.
×
Page 326

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

CHAPTER VII Resources for Basic Research in the Chemical Sciences INTRODUCTION The understandings that follow from basic research in chemistry open new options for addressing societal needs. The benefits may be recognized and realized within only a few years, or they might not come to fruition until decades after the most crucial discoveries. This range of time horizons explains, in part, why research in chemistry is conducted in the United States in various arenas, industrial laboratories, private (not-for-profit) laboratories, national and other federal laboratories, and in our university and college laboratories. Progressively through this sequence, research tends to be increasingly directed toward the fundamental understanding of nature and less practical or goal- oriented. This trend reveals the fact that in the United States, the frontiers of chemistry are primarily explored and expanded in our research universities. This characteristic contrasts with the organization of fundamental research abroad, as witnessed by the considerable dependence upon the Max Planck Institutes in Germany, the CARS Laboratories in France, and the Academy Institutes in the Soviet Union. The U.S. system has the enormous advantage that it strongly couples the basic research function to the education of the next generation of scientists. Thus it continuously renews our pool of scientific personnel with young scientists whose thesis work has probed the edges of our knowledge. BASIC CHEMICAL RESEARCH IN INDUSTRY Because of the clear potentiality for short-term payofffrom chemical research, the chemical industry invests heavily in its own in-house research. In 1982, the Chemical and Allied Products industries invested about $4.2B in corporate research and development, of which about $380M might be classified as basic research. These statistics implicitly confirm a major thesis of this report, that 288

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES research in chemistry pays oh in future processes and products used by society. They also show that industrial laboratories furnish an important locus of chemical research. Excellent fundamental research is conducted in this arena but, by and large, attention tends to be focussed on programs that offer prospect for marketable products within a fairly short time, and most of these programs must be proprietary. Scientists in industrial laboratories depend upon and draw from a reservoir ot fundamental knowledge constantly renewed and expanded by university-based research. Industries also rely upon a stream of talented young scientists entering the field and bringing immediate familiarity with the latest discover- ies, the recent scientific literature, and the newest instruments and scientific techniques. As tangible evidence of industry's recognition of these dependen- cies, the chemical industries furnish direct support to university research in amounts estimated to be about $10M to chemistry departments and $10M to departments of chemical engineering in 1983. While this is only a modest fraction of the resources needed to maintain international research leadership for U.S. chemistry, it is an extremely important source of support. Industrial support of university research provides communication and coupling between industrial and academic scientists that facilitate movement of new discoveries into the industrial laboratories where applications can be developed. At the same time it can influence beneficially the graduate educational process, and it gives industry some opportunity to influence the university research agenda. Efforts are being made to strengthen this coupling (e.g., through the Council for Chemical Research) and to increase the amount of this support. A realistic appraisal suggests that industrial support might be as much as double its current amount. Tax incentives to encourage such gains should be explored. It is recommender! that new mechanisms and new incentives be sought for developing stronger links between industrial and academic research and for increasing industrial support for fundamental chemical sciences research conducted in universities. The most fundamental and adventurous research will remain a modest, though vital, component of industrial research because the likelihood for payoff is too uncertain and the time horizon for application is too remote. Yet, the most fundamental chemical discoveries can offer the most far-reaching benefits to society, most often in directions that could not have been foreseen. This "high-risk," "blue sky" research ultimately furnishes the intellectual basis for our cultural ethos and our technological competitiveness. Hence it is an appropriate place for public investment. it explains and justifies the consider- able federal investment in support of scientific research at the national laboratories and at the nation's university laboratories. Total federal obligations for basic research in chemistry were $349M in fiscal year 1983. When corrected for inflation, this represents a 10.9 percent increase over the federal investment 10 years earlier. During that same period, the sum 289

290 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES (corrected for inflation to 1972 S's) - c~ 0$100B - - He S $50B o PETROLEUM and COAL /~< - 1 Ul1LM 1UAL PRODUCTS MACH I NERY D/7 \~` MOTOR ...... VEH ICLES 972 1 974 1 976 1 978 1 980 1 982 YEA R REAL GROWTH FOR CHEMISTRY-BASED INDUSTRIES of the chemistry-based petro- leum, coal, and chemical industries increased in in- flation-corrected business vol- ume by more than a factor of two, while machinery deliver- ies were increasing by only 30 percent and automobile deliv- eries were 20 percent below 1972 levels. (All figures cor- rected for inflation.) Equally important, the in- ternational balance of trade for chemical products has steadily remained positive. It has risen from $1.4B in 1965 to about $12B in 1981, second only to machinery. Thus our chemistry-based industries are key to our overall eco- nomic well-being, and their future must be assured. While the FY 1983 federal obligation of $349M for chemistry research may seem a large sum, it is well within bounds defined by federal obligations for other physical sciences that depend upon sophisticated instrumentation (see Table VIl-1), which compares federal funding for chemistry novice and astronomy over the 10-year period 1973 to 1983. ~ 7 ~——A/ __ _~ ~ The parenthetical numbers, corrected with GNP deflators to 1973 dollars, show that while real growth in chemistry funding was Il percent, physics funding grew 20 percent and astronomy funding 44 percent. At least one criterion by which these figures can be judged is their appropriateness to the need for talented young people in the nation's industries. Thus if one divides the federal obligations (in FY 1983 dollars) by the number of scientists employed in industry, this "normalized" annual investment in chemistry amounts to $4K per year per employed chemist. This is 16-fold less than the comparable investment in physics and astronomy (taken together because there are no separate data on the use of astronomers in industry). A related "normalization" can be based on the number of Ph.D. degrees granted per year. In 1983, about 1700 chemistry Ph.D. degrees were conferred, so the annual investment per Ph.D. joining the work force was $205K. This figure is one-fifth the comparable figure for physics and perhaps one-twentieth that for astronomy. Thus, the fecleral investment in chemistry is meagre comparer! to that receiver! by its companion physical sciences. Cllearly, this investment is not commensurate

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES TABLE VII-1 Federal Obligations for Basic Research in the Physical Sciences, 1973 and 1983 Chemistry Current $ (FY 1973 $)a $146M (146M) $349M (162M) Physics Current $ (FY 1973 $)a $351M (351M) $905M (421M) Astronomy Current $ (FY 1973 $)a $122M (122M) $379M (176M) (1) FY 1973b (2) FY 1983b (3) ScientistsC employed by industry 86,600 (1980) (4) Number Ph.Ds~ (1983) $/Industry scientists (2)/(3) $/Ph.D. (2)/(4) 1700 $4.0K $205K 22,400 830 100 $57.3K $1090K $380K a Based on GNP deflator, see Science Indicators, 1980. b See Table A-7. c See Table A-4. ~ See Table VI-1. with the practical importance of chemistry both economic and societal nor with the outstanding intellectual opportunities it now offers. To frame recommendations directed toward redress of this imbalance, we must examine the budgets of each of the federal agencies that might logically support basic chemical research in the public interest and in the achievement of their respective missions. Table VIT-2 shows the distribution of support for chemistry among the five agencies that contribute significantly to fundamental research performed in universities and colleges. These agencies also support research in their own national laboratories, but funding of basic research performed in universities and colleges is a clear-cut and unambiguous indicator of a particular agency's commitment to long-range chemistry research and the renewal of the pool of scientists. Table VIl-2 shows that the largest fraction for the support of chemical research has come from the National Science Foundation over the last decade. This is despite the obvious relevance of chemistry to the congressionally mandated missions of the other agencies listed. The detailed budgets of each of these agencies will be considered in turn. First, however, it is appropriate to consider the "style" of research in chemistry. CHEMISTRY: AN ACTIVITY OF CREATIVE INDIVIDUALISTS Today's public image of science is still heavily influenced by the reverberating impact of the World War II Manhattan Project that brought us the atomic bomb and the Apollo Project of the 1960s that let us set foot on the Moon. We are seen to be in an era of Big Science. But embedded in this glamorous, highly organized, and well publicized setting, there are a number of scientific disci- plines that have somehow maintained the highly personal characteristics of 291

292 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES TABLE VII-2 Federal Obligations for Basic Research in Chemistry Performed by Universities and Colleges (by Percent of Total) NSF NIH DOD DOE DOA FY 1974 49.2 FY 1984 44.7 24.2 21.5 9.2 9.2 16.1 11.0 3.1 3.6 classical human creativity: How many writers were needed to create Hamlet? How many artists to paint the Mona Lisa? How many scien- tists to propose relativity? Chemistry is one of these dis- ciplines. Somehow it has re- mained an idiosyncratic and highly competitive activity that depends upon ~ i, ~ . sustained individual initiative and personal creativity. Scientific publications in the field generally involve two or three authors. There are no examples to be found in chemistry to match the multiple authorship—dozens of authors on a single paper—like those announcing the occasional discovery of a new subatomic particle. Chemistry has remained, worldwide, an innovative "cottage industry" with a modus operandi that has been remarkably productive. Tangible evidence of its success is provided by the faster-than-exponential discovery of new compounds (see Chapter I, p. 41. This gratifying record was achieved despite the fact that at any given moment, the molecules easiest to synthesize have already been made; the harder ones remain. Yet discovery is accelerating. The only plausible explanation is that chemistry in the small project mode is an extremely effective enterprise, both here and abroad. Thus the term "cottage industry" describes a highly individualistic and personally creative activity rather than a consensual one. These characteristics impart a healthy competitiveness and a liberating freedom from bounding paradigms. They make chemistry an ideal field in which to nurture a young scientist's originality and initiative. He or she can be intimately involved and in control of every aspect of an investigation, selecting the question, deciding on the approach, assembling and personally operating the equipment, collecting and analyzing the data, and deciding on the significance of the results. Here is another reason to nourish this central and fundamental science in its present image. Yet we are in an era in which directors of U.S. federal science-funding agencies will candidly admit that they believe it easier to argue for an enormous increment of funding to sponsor a large machine or a massive project than for a smaller increment to stimulate many smaller projects with comparable or greater expectation for new discoveries and scientific advances that will surely respond to society's needs. Thus the Department of Energy in its 1985 budget devotes 55 percent of its Office of Energy Research budget to two "Big Science" project areas: $54SM for high energy physics and $440M for fusion research. Currently under consideration by various agencies are proposals for incremen- tal funds to build a hard X-ray synchrotron light source ($160M), a neutron source ($250M), a "next-generation" multimirror telescope ($100M), a set of "supercomputer centers," an array of"engineering centers," and an accelerator

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES with a circumference of 80 miles (CRIB). Each of these new projects will require large, ongoing (and incremental) operating budgets that are irresistibly rooted in huge initial capital investments. In the presence of such ambitious programs, the incremental resources needed to exploit the rich opportunities before us in chemistry are easily in scale. Because of the societal payoff to be expected from such an incremental investment, it will be readily and persuasively defensible in the individual competitive grant style already known to be elective in chemistry. PRIORITY AREAS OF CHEMISTRY The strength of American science has been built on allowing creative, working scientists to decide independently where the best prospects lie for significant new knowledge. Many of the most far-reaching developments, both in concept and application, have come from unexpected (Erections. Thus, a listing of priority areas may tend to close ok or quench some of the most adventurous new directions whose potential is not yet recognized. Even so, it makes sense to concentrate some resources in specially promising areas. This can be done if we regard our research support as an investment portfolio designed to achieve maximum gain. A significant part of this invest- ment should be directed toward consensually recognized priority areas but with a flexibility that encourages these favored listings to evolve as new frontiers emerge. A second substantial element in this portfolio should be directed toward creative scientists who propose to explore new directions and ideas. Then, a third element must be the essential resources to provide the needed instrumen- tation and the infrastructure for its cost-e~ective use in achieving the goals of the entire portfolio. Where this balance will fall for each of the funding sources will vary, of course. Industrial research will weight rather heavily the currently recognized priority frontiers. At the other extreme, NSF must take as its first responsibility the encouragement of new avenues from which tomorrow's priority lists will be drawn. The other mission agencies should structure their portfolios between. This report shows decisively that this is a time of special opportunity for intellectual advances in chemistry. Furthermore, the report demonstrates that such advances will not only enrich our cultural heritage, but also will help us respond to human needs and sustain our economic competitiveness. It is in society's interest to exploit these opportunities and to do so with particular attention to those frontiers that deserve high priority because they can be confidently expected to yield high intellectual and social return from the needed additional federal investment. We identify here five areas that meet this criterion. A. Understanding Chemical Reactivity B. Chemical Catalysis C. Chemistry of Life Processes 293

294 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES D. Chemistry Around Us E. Chemical Behavior Under Extreme Conditions Understanding Chemical Reactivity This is surely a time of special opportunity to deepen our fundamental knowledge of why and how chemical changes take place. The advance of the frontiers of reaction dynamics at the molecular level has undergone a revolu- tionary advance during the last decade. At the same time, synthetic chemists are constantly adding to our arsenal of reaction types and classes of compounds in a way that is eliminating historical distinctions between organic and inorganic chemistry. Much of this remarkable progress is due to the develop- ment and application of powerful instrumental and analytical techniques that give us capability to probe far beyond current bounds of knowledge. In reaction dynamics, we can now aspire to elucidate the entire course of chemical reactions, including the unstable structural arrangements interven- ing between reactants and products. Just as the last three decades saw rich development of our understandings of equilibrium molecular structures and equilibrium chemical thermodynamics, the next three decades will see elucida- tion of the temporal aspects of chemical change. We will be able to ascertain the factors that determine the rates of chemical reactions because of our new abilities to watch the fastest chemical processes in real time, to conduct reliable theoretical calculations of reaction surfaces, to examine chemical changes at the most intimate level ("state-to-state"), to track energy movement within and between molecules, and to exploit hitherto inaccessible nonlinear photon excitation processes ("multi-photon" excitation). These remarkable possibilities are rooted in a powerful array of new instruments, foremost of which are lasers and computers, and including Fourier transform infrared spectrometers, ion cyclotron resonance techniques, molecular beams, and synchrotron radiation sources. New reaction pathways in synthetic chemistry offer another rapidly advancing frontier. These pathways identify a high leverage opportunity because they provide the foundation for future development of new products and new pro- cesses. Selectivity, the key challenge in chemical synthesis, is the cornerstone. Control of the different intrinsic reactivity in each bond type (chemoselectivity), the connection of reactant molecules in proper orientation (regioselectivity) and in the desired three-dimensional spatial relations (stereoselectivity) is at last within reach. Our ability to produce a controlled molecular topography has far-reaching implications for catalyst design. The traditional line of demarka- tion between organic and inorganic chemists has virtually disappeared as the list of fascinating metal-organic compounds continues to grow. We have just begun to elaborate and understand the potentialities of chemical pathways opened using light as a reagent. Finally, chemists are learning how to prepare solids with a wide range of tailored properties that include inorganic solids with contrived cavities as designed catalysts, polymers with structural properties

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES that challenge those of steel, and new families of"electronic chemicals"- inorganic and organic semiconductors, resists, super-lattice materials, optical fibers, nonlinear optical materials that will accelerate development of micro- electronics and information transition. Again powerful instrumentation plays a central role. Rapid and definitive identification of reaction products, both in composition and structure, account for the speed with which synthetic chemists are able to test and develop adventurous synthetic strategies. Of prime importance are high-resolution Nuclear Magnetic Resonance, computer-controlled X-ray crystallography, and high-resolution mass spectroscopy coupled with the delicate separation capabil- ities provided by chromatography in its advanced forms. Chemical Catalysis A catalyst accelerates a chemical reaction toward equilibrium without being consumed. This acceleration can be as much as 10 orders of magnitude while favoring one particular reaction out of many competing pathways. There is now in prospect the possibility of obtaining a molecular-level understanding of catalysis to move it from an art to a science. Rich payoff can be expected because we will be laying the foundation for the development of new technologies. All facets of this critical frontier are opening and synergistically interacting. Heterogeneous catalysts are solid materials prepared with large surface areas upon which chemical reactions occur at extremely high rate and selectivity. Entirely new kinds of information are now accessible through an arsenal of new detection techniques of such sensitivity that we can hope to watch chemical change take place on a solid surface flow energy electron diffraction, electron energy-Ioss spectroscopy, Auger spectroscopy, photoelectron spectroscopy, sur- face-enhanced Raman, etc.~. These instruments open the door to understanding and controlling the chemistry in this surface domain. Homogeneous catalysts are soluble and active in a liquid reaction medium. Often they are complex metal-containing molecules whose structures can be modified to tune reactivity in desired directions to achieve high selectivities. Organometallic chemistry and metal cluster compounds are of particular importance; they reveal homogeneous catalysis as a bridge between heteroge- neous catalysis and enzyme catalysis. Artificial enzyme catalysis is now an exciting frontier because of our instru- mental capability to deal with molecular systems of extreme complexity. It permits us to apply the synthetic chemist's ability to produce a molecular topography designed to adsorb selectively a paticular reactant and hold it in a known geometry the enzyme counterpart to the basic feature of a heteroge- neous catalyst. At this structured surface locale, a chemically bound metal atom is placed so that it will impart to the adsorbed guest molecule a desired chemical reactivity—the enzyme counterpart to homogeneous catalysis. Prototype arti- ficial enzymes are in preparation and are on the drawing boards. Revolutionary potentialities for new processes are in prospect. 295

296 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES Electrocatalysis and photocatalysis are adding new dimensions to the field of catalysis. Chemical reactions can now be induced at the interface between a liquid solution and an electrochemical electrode surface, with or without absorption of light by a semiconductor used as an electrode. In either case, our growing knowledge of homogeneous catalysis and of semiconductor behavior is being applied and coupled with the stimulating aspect of the electro- or photochemical energy input. Potential applications range from solar energy storage to photogeneration of liquid fuels, such as methanol from carbon dioxide and water. Chemistry of Life Processes In the last decade, chemists have succeeded in recognizing and synthesizing a large number of molecules of exquisite complexity. This capability is most timely because the exciting and largely phenomenological advances of the biosciences now demand explication at the molecular level. Thus, the time is ripe for quite spectacular advances in chemistry at the interface of chemistry and biology. These advances are bound to have applications to human health, animal health, and agriculture. The opportunities are illustrated by three broad types of problems: receptor-substrate interactions, vectorial chemistry at mem- branes, and genetic engineering. Receptor-substrate interactions selectively mediate essentially all biological processes. Thus, protein receptors (enzymes, antibody, membrane, or intracel- Jular receptors) interact selectively with one or more substrates (enzyme substrate, antigen, hormone, neurotransmitter, or simple molecule or ion). Chemistry is needed to understand these processes in molecular detail because we must be able to isolate and identify the structures of these substrates, synthesize them in useful quantities, analyze their receptor-interactions in physical-chemical as well as biological terms, and modify their structures to suit desired uses. Medical and biological applications of great value can be foreseen, but the methods of chemistry are needed in order to manipulate these substances. Vectorial chemistry describes reactions depending upon spatial separation (as by a membrane) of reactants into regions of different concentrations. Because of the importance of such systems in living systems (celIs and organelles), we must understand the relations between concentration gradients across membranes and the processes by which chemical bonds are made and broken. Active chemical modelling, at both the synthetic and mechanistic levels, is needed to complement current activity in mechanistic biology. Progress will be directly applicable in pharmacokinetics and drug delivery. Genetic engineering has become possible as molecular biologists have discov- ered and exploited the action of certain natural enzymes that affect DNA. Restriction enzymes catalyze the cutting of DNA at special places, and ligation enzymes can join it (or a contrived insert) together again. Through synthesis of DNA and RNA sequences, structural analysis of gene fragments, and develop-

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES ment of separations techniques, chemists will play an important and increasing role as we use these capabilities to clarify, on a molecular level, the chemistry of genetics, and as we add to the growing number of applications of genetic . . engineering. Chemistry Around Us The atmosphere, oceans, earth, and biosphere are strongly coupled through the chemical processes that take place in each region. Man's disturbance of this global chemical reactor is no longer negligible. Networks of interlocking chemical cycles involving trace constituents help determine the gross structure and behavior of the stratosphere, the troposphere, and, through rain, the soil and lakes that make up our local environment. To protect these local environ- ments, we must understand what chemical substances are present and in what concentrations, as well as what chemical reactions they induce, and at what rates the changes take place. The first two issues involve analytical chemistry, and the second two involve reaction dynamics. Fortunately, both fields are in particularly fruitful states of development. Analytical chemistry critically determines our ability to advance our under- standing of environmental chemistry because much of this chemistry is con- trolled by reactive molecules present at trace levels in some cases as low as parts per trillion. Astonishing progress is being made in extending analytical sensitivity limits (How little can we detect?), sharpening analytical specificity (How sure can we be of what we are detecting?), and improving separations (Can we isolate the desired constituent even in miniscule quantities?. Such progress has immediate applicability in the analysis of complex but very dilute mixtures of pollutants, pesticides, and degradation products of both human and natural origin as found in ambient air, toxic wastes, polluted streams and lakes, agricultural soils, and biological samples. Instrumentation will play a central role in these gains. Analytical chemists are applying our most sophisticated techniques, including tandem mass spec- trometry, high-resolution gas chromatography coupled with mass or Fourier Transform infrared spectroscopy, supercritical fluid chromatography, remote detection using laser fluorescence or absorption techniques, ultrasensitive in-cavity and photoacoustic laser methods, chemiTuminescence, and computer- aided data collection and manipulation. Increased investment in analytical chemistry will permit us to extend detection well below toxicity bounds so that potential problems can be anticipated and ameliorated Tong before the hazard level is reached. Reaction dynamics in environmental chemistry poses some surprisingly difficult problems that define new research fronts. In atmospheric chemistry, complex chains of interlocking reactions can be involved, and reaction rates of highly reactive and elusive molecules can be controlling (e.g., OH, HO2, NOT. Because of the reactivity, these rates are difficult to measure, and reliable laboratory methods for measuring the relevant rate constants must be devel- 297

298 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES oped. Even the locale of reaction may be in question because of the presence of solid and liquid particulates, including finely divided carbon particles (soot). Reaction rates here may be catalytically enhanced. Nucleation of aqueous clusters induced by such pollutants as sulfur oxides and nitric acid may figure importantly in transport as well as chemistry, so nucleation rates must be known. Hypersaline droplets provide an unfamiliar reaction environment in which aqueous solution chemistry can be strongly perturbed. Photochemistry is a significant factor and adds to the problem because the chemical fate of the photoactive species must be considered in all the potential reaction locales in the gas phase, adsorbed onto solid particles, or dissolved in droplets (hypersaline or not). For example, little is known about aqueous phase photo- chemistry in clouds. Once again, modern laser spectroscopic techniques and computer-aided data collection, usually coupled with traditional kinetic methods, are permitting us to address these challenging questions. We cannot afford to neglect the rnnctinn dynamics aspect of environmental chemistry. Chemical Behavior Under Extreme Conditions Most of our knowledge of chemical change has been accumulated within a narrow range of the influential variables, pressure and temperature being the most obvious. Now, as our measurement techniques are becoming more power- ful, we can investigate chemical processes as they occur under conditions far removed from those of our normal ambient surroundings. The ability to study chemistry under such extreme conditions expands the number of laboratory variables with which chemical reactivity can be manipulated and controlled. At the same time, these extreme conditions provide critical tests of our basic understanding of chemical processes. With capabilities in hand or close on the horizon, significant progress in new materials, new processes, new devices, and deeper understandings would reward a concentration of effort on the study of chemistry under extreme conditions. The effort should encompass chemistry under exceptionally high pressure, high temperature ,, in gaseous discharges plasmas J. and at temperatures near absolute zero. High-pressure chemistry has potentiality on several fronts as it has become possible to examine reactivity at pressures up to and exceeding a million times atmospheric pressure (>1 megabar). High-pressure studies of reactivity reveal the volume profile of reactants so they add an entirely new facet to our description and understanding of the unstable atomic arrangements that intervene between reactants and products. The insights so gained could be one of the important ways that the temporal aspects of reactions (reaction rates) are understood and brought under control. Reaction mechanisms are revealed, and reaction pathways can be manipulated. New processes and chemical products are to be expected. Next, pressure can have a differential effect on electronically excited states of molecules, thus altering the optical properties of liquids and solids, and it can

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES change the packing of molecules in the solid state to affect electrical properties. The ability to vary the relative energies of electronic states in a continuous manner can provide unique tests of theories of electronic processes, energy transfer processes in phosphors, and behavior of laser materials, such as the IT-VI and IlI-V compounds. It has already been shown to provide control of photochromism, semiconductor-conductor transitions, and superconducting transitions. Further gains in these areas can be expected to result in new devices, such as pressure-tunable display devices, lasers, switches, superconduc- tors, and magnetic devices. Thirdly, we are seeing a revolution in our understanding of critical phenom- ena, the behavior of substances at the pressures and temperatures at which the gaseous and liquid states can no longer be distinguished. Powerful new theoretical approaches are responsible, such as the "renormalization group" approach. A further unexpected discovery is that surfaces and interfaces of near-critical fluids can themselves exhibit transistions and critical phenomena. Experimental techniques are keeping pace, as revealed by both prediction and demonstration of"wetting transitions." Applications are in store, such as microemuIsions for of] extraction, liquid crystals for display devices, and new polymer transitions. High temperature chemistry is not new: combustion has been known since prehistoric times. What is new is our ability to sustain temperatures up to 6000 K over large volumes for long times and to produce temperatures very much higher (up to 100,000 K) for shorter times through a variety of pulsed techniques (laser pulses, shock waves, explosions, and electrical discharges). Exploiting these possibilities will result in more efficient use of energy, high temperature materials, high temperature fabrication processes, and new un- derstandings of chemical reactivity, including combustion. High temperature liquid phases are present in many technological processes (molten metals, silicates, salts, oxides, sulfides, metalloids), but fundamental understandings have lagged those for gases and solids. Unusually strong and sometimes long-range interactions characterize these liquids, and techniques for their study (including pulsed laser, X-ray, neutron diffraction, Raman scattering, NMR and EXAFS) must be adapted to the high temperature regime. The technological importance of high temperature solids is already obvious and is still increasing. Substitutes made necessary by scarcity (e.g., for chro- mium alloys) or environmental problems (e.g., for asbestos) and materials that must function at extremely high temperatures (e.g., in gas turbines, nose cones, internal combustion engines) provide examples where substantial gains of practical importance are to be expected. Finally, the basic foundations for prediction are woefully inadequate at high temperatures where gross excitations of electronic, vibrational, and rotational degrees of freedom make first-order simplifying assumptions (such as separa- tion of these variables) inadequate or inapplicable. Spectroscopic data are needed for high temperature species, both in ground and excited electronic 299

300 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES states but with high vibrational and rotational excitation. Bond dissociation energies, heats of formation, ionization energies, and radiation cross sections must be reliably known for the unfamiliar molecular species found in flames and explosions. The applicable thermochemical data base must be filled in and expanded. Plasmas are produced by an electric discharge through a gas. Extremely high temperatures are characteristic; they can reach a million degrees Kelvin. A nonequilibrium situation can be expected, and the apparent temperature can depend upon the species and the degree of freedom probed. Radiation within and out of the plasma zone can contribute to the behavior. The chemical species present include ions, electrons, and electronically excited and chemically reactive neutrals. These characteristics give plasmas unique and, at this time, poorly known chemistry, so much so that plasmas have been called a new state of matter. Better understanding of plasmas will be of value because of their growing importance in a wide variety of contexts ranging from stellar chemistry to practical applications in semiconductor fabrication (plasma etching) and design of nuclear fusion reactors. Low temperature chemistry, once thought to be nonexistent, is providing unique information about chemical reactions that take place at temperatures near absolute zero. Such reactions can now be studied both in the gas phase, using supersonic jet cooling, and in solid environments through the matrix isolation method. The supersonic jet cooling technique has helped open the spectroscopic study of weakly bound molecules ("van der Waals molecules". Coupled with laser-induced fluorescence and molecular beams, it opens new avenues to understanding intramolecular energy transfer and predissociation. Also, the ability to prepare complex molecules in the gas phase but with vibrational and rotational temperatures near absolute zero gives access to spectroscopic detail normally completely obscured. New applications of super- sonic jet cooling are still being discovered, and they often provide unique insights. Suspension of highly reactive molecules in inert gas solids at temperatures near absolute zero (the matrix isolation technique) also yields information difficult to obtain by any other method. The results nicely complement both high temperature studies and supersonic jet cooling because the structures and reactions of the transient molecules that determine the chemistry of flames and explosions can be examined. Recent applications have turned to unusual reactions, such as metal atom insertions into organic molecules and light- induced isomerizations. Laser-selective excitation of cryogenically trapped reactive pairs has provided the first evidence of mode-selectivity in bimolecular reactions, and it has suggested new chemical routes to solar energy storage. Chemical synthesis at extremely Tow temperatures has also provided prepar- ative routes to molecular species too reactive to accumulate under normal conditions. A number of cryogenic solvents permit synthesis of complex organic and metal-organic molecules at temperatures reaching down to that of liquid

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES air. Even inert gases can be liquified to use as reaction solvents over a wide range of cryogenic temperatures, using high pressure to broaden the liquid ranges. Finally, evaporation and sputtering of metal atoms into cryogenic organic samples can be used in large-scare preparative processes for unusual metal-organic substances that may have special catalytic activity. CHEMISTRY AND THE NSF MISSION The chemistry supported by NSF is judged mainly on its potential for adding to our understanding of nature rather than its presumed relevance to a practical outcome. Because the most far-reaching technological changes tend to stem from unexpected and unpredictable discoveries, the fundamental research supported by NSF is most critical to the long-range technological future of this country. In NSF, the chemical sciences are supported primarily in three divisions: the Chemistry Division, the Materials Sciences Division, and the Chemical and Process Engineering Division. Chemists also participate to some extent in divisional programs in other NSF directorates: Biochemistry, Atmospheric Chemistry, Geochemistry, and Marine Chemistry. The Materials Sciences Division includes three chemical science-oriented programs: Solid State Chem- istry, Polymers, and Metallurgy. Also included are the Materials Research Laboratories (the MRL Program), which block-fund groups of physicists, chem- ists, and engineers with relatively focussed and applied program goals. Table VIl-3 shows the budgets for these programs over the years 1972 to 1984, both in current dollars and corrected with GNP deflators to constant, 1972 dollars. Table VIl-3 shows that over the decade 1972-1983, total NSF funding for the chemical sciences rose in real terms by about 40 percent; but this growth was concentrated in the materials sciences, which grew by 50 percent and in chemical engineering, which grew from a small base by about 90 percent. The Chemistry Division budget grew by 23 percent over the same period, during which costs of scientific instrumentation have increased much faster than inflation, and day-to-day dependence upon such instrumentation has become almost universal across chemistry. Over this same decade, our economy has depended more and more on the health of the chemical industry, as reflected in both the doubling of its contribution to the GNP and the growth of its positive trade balance. These factors, coupled with the exciting opportunities in chem- istry, justify considerably larger NSF support. To assess the extent to which incremental funds are needed, it is necessary to examine in detail the disposition that has been made of the available resources. This can be examined most clearly for the Chemistry Division budget and in terms of three funding dimensions: purchase of large instrumentation for shared use (either on a departmental or regional basis), purchases of expensive instrumentation for dedicated use, and the size of individual research grants. To 301

302 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES TABLE VII-3 NSF Funding of the Chemical Sciences, 1972-1984 Materials Materials Chemistry Science Chemistry Science Division Divisiona Chemical Division Divisiona Chemical Year (in current-year dollars) Engineering (in constant 1972 dollars) Engineering 1972 24.5M 4.8M 24.5M 4.8M 1973 25.1 5.0 24.0 4.8 1974 26.6 5.1 23.7 4.6 1975 32.7 5.8 26.5 4.7 1976 34.7 8.2 26.3 6.2 1977 40.2 9.8 28.6 7.0 1978 43.1 11.3 28.7 7.5 1979 45.2 11.5 13.0M 27.6 7.0 8.0M 1980 51.7 14.7 13.9 29.1 8.3 7.8 1981 57.6 14.9 16.7 7.6 8.6 1982 61.4 15.9 20.4 29.4 7.6 9.8 1983 67.6 16.2 22.6 30.8 7.4 10.3 1984 (est.) 80.0 19.5 26.6 33.2 8.1 11.1 1985 (request) 92.1 22.6 32.5 35.6 8.7 12.5 1985 (est.)b 87.8 21.5 33.8 8.3 a This is the portion of the MRL budgets received by members of Chemistry and Chemical Engineering Departments, about 20 percent of the total. The remainder is directed to members of Physics and Engineering Departments who receive, respectively, about 45 percent and 35 percent of the total. b After congressional action, as of January 1985. begin, it is helpful to trace the history of funding by the Chemistry Division for these three elements. Before 1970, instrumentation accounted for less than 7 percent of the Chemistry Division budget. It was then deliberately raised to about 9 percent NSF CHEMISTRY DIVISION BUDGET for the years 1970 to 1972. Again a deliberate decision in 1973 led to another increase in the percent of the Chemis- try Division budget spent for dedicated and shared instru- mentation, this time to 17 percent. This percentage slowly rose to 18.6 percent by 1982. Now, in 1984 and 1985 it is about 25 percent. Through much of this 15- year period, the amount ex- pended for dedicated instru- mentation has been essen- tially constant at about 10.5 percent. The major chang- z o Z 20 z ~ a: o en z x TOTAL INSTRUMENTATION 1 _ I fry SHARED INSTRUMENTATION | DEDICATED 10 1 970 1 975 1 980 1 985 YEA R STEADILY RISING INSTRUMENTATION COSTS ( IN PERCENT OF CHEMISTRY D IY ISION BUDGET)

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES es have been in response to the pressure for heavier investment in shared instrumentation. Taken together, shared and dedicated instrumentation have eaten away a larger and larger fraction of the Chemistry Division budget with the result that almost no real growth could be directed toward the other costs of performing research. Over the 11-year period 1972 to 1983, the budget less instrumentation rose only ~ percent (corrected for inflation). The need to divert to instrumentation an increasingly larger fraction of the NSF Chemistry Division budget is but one of the symptoms of resources inadequate to the opportunities we must exploit and develop. To remedy this, we present an analysis of the needed level of resources in each of the major dimensions: shared instrumentation, dedicated instrumentation, and individ- ual grant size. The analysis is based on the following four premises, the first of which is clearly established throughout this report. Chemistry is rich with opportunities for intellectual advances, for improve- ment of the quality of life, and for strengthening our economic competitiveness. We wish to assure a U.S. position of international leadership in the exploitation and development of these opportunities. We wish to attract to the chemical sciences an appropriate share of most promising young scientists to provide the talented manpower needed in this vital area. We wish to be sure that this enterprise engages the full national capability, as is appropriate to the national benefits to be derived. Shared Instrumentation The NSF has a proud heritage in its support of departmentally or regionally shared instrumentation. Unfortunately, through the 1970s, the NSF carried most of the load in supplying and maintaining such equipment in chemistry. However, current resources are quite insufficient. Many worthy requests must be declined, up to 50 percent cost-sharing might be required, and no allowance is made for ongoing maintenance. With the above four premises in mind, we offer here a systematic approach to estimating the need. The NSF shared instrumentation program receives requests for a variety of instruments in the cost range of $200K to $800K. These requests reveal the instrumental techniques needed to conduct frontier research in chemistry. Table VIl-4 lists eight examples of costly instrumentation commonly requested. No claim is made that every chemistry department needs every one of these instruments nor even a particular subset of them. We do assert, however, that a research university must maintain state-of-the-art capability for some of these instruments to remain front rank. The breadth of the research strength for a particular department will determine how many instruments would be needed to maintain international competitiveness. A large department with full breadth would require six state-of-the-art instruments in its "Departmental Instrument Inventory." With a somewhat narrower range of activity, three 303

304 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES TABLE VII-4 in Shared Instrumentation International Competitiveness Departmental instrumentation menu (1) High-field NMR spectrometer (1) Mass spectrometer (1) X-ray spectrometer (1) FTIR (1) Raman (1) ESR (1) Departmental computer (1) Array processor (8) Instruments Average capital cost per instrument Cost share, NSF, 80~o Maintenance and operation at 20~o capital cost/year for 5 years 750K 500K 450K 200K 200K 200K 300K 250K $2850K 355K 285K Total cost, NSF share per instrument $640K instruments might suffice. With a still sharper research focus, two instruments would probably keep a number of research faculty active at their research frontiers. Thus we begin this analysis by con- sidering three "Departmental Instrument Inventory" lev- els six, three, or two state- of-the-art instruments from the list in Table VIl-4. At each level, the instruments must be replaced often enough to avoid obsolescence. $355K Experience in industrial re- search laboratories indicates that replacement time for the modern instrumentation un- der consideration averages about 6 years, a time determined by technical, not mechanical obsolescence. In other words, well before instrument-maintenance becomes a major problem, new instrument capabilities have become available that redefine the feasible research goals. On the average then, over a 6-year period, a department must have a reasonable chance to compete for continued access to the largest state-of-the-art capability for its Departmental Inventory. The financing of such a program includes three elements: initial capital investment, cost-sharing policy, and ongoing costs of maintenance and opera- tion. It is time to carry over from physics and astronomy the well-recognized principle that cost-effective use of sophisticated instrumentation requires ongo- ing support for maintenance and operation. Without this infrastructure sup- port, such instrumentation is used neither at full efficiency nor full effective- ness. On the basis of studies on this issue, about 20 percent of the original capital investment for a 5-year period is needed. NSF should build into its shared instrumentation program annual operating and maintenance support (20 percent of capital investment) for a period of 5 years after purchase. Again we learn from physics and astronomy that cost-sharing by the institu- tion becomes less and less viable as instrument costs continue to rise. Given the current financial stress in universities, we urge that the federal share of instrumentation capital investment should average at 80 percent. With these guiding recommendations on cost-sharing and infrastructure support, Table VIl-4 shows that an average cost per shared instrument would require an NSF commitment of $640K. Finally, we propose a deliberate approach to the question of how many U.S.

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES research universities can and ought to be In mind as NSF sponsors academic research in chemistry. There are 198 U.S. institutions that grant the Ph.D. degree in chemistry. The corresponding graduate faculty exceeds 4000, the number of postdoctorals is about 2800, and the number of full-time graduate (M.S. and Ph.D.) students exceeds 13,000. Of the 198 degree-granting institu- tions, we propose that at least 20 of them should be able to compete successfully for a Departmental Instrumentation Inventory of six major instruments. Such an institution could consider replacement of each of these instruments every 6 years, to average one per year. For a larger number of institutions we propose 40 that competition might be successful every other year. For these, the Departmental Inventory could include three state-of-the-art instruments. For a next tier of 60 institutions, success would average 3 years, to maintain two modern instruments. These estimates 20 first-tier, 40 second-tier, and 60 third-tier institutions~an be seen to be modest in scale by comparing to the number of Max Planck Institutes in Germany. There are a dozen Max Planck Institutes in which chemistry is a central component. These Institutes might be considered analogous to our 20 first-tier research universities because they tend to be more richly endowed with instrumentation than the larger number of German universities also engaged in chemistry research. Scaling on the basis of total population, a dozen Institutes in Germany would correspond to about 40 competititive institutions in the United States (3.4 times higher). This shows that the level we propose is reasonable and needed if we wish to maintain international competitiveness. The cost of such a shared instrumentation program is shown in Table VIT-5 and compared with the projected NSF Departmental Instrumentation budget for FY 1985. The NSF shared instru- mentation budget falls far short of the need. The NSF Departmental Instru- mentation budget for FY 1986 should be augmented by $25M. Dedicated Instrumentation Much fundamental research today depends upon the sophisticated instrumen- tation just discussed but with such specialized application that shared use is quite impractical. However, the philosophy used above, based upon the same four premises, again fur- nishes an estimate of the re- sources needed to maintain U.S. leadership in a number of crucial specialty areas. This time an average cost per instrument is based upon a 1 5-instrument menu that weights some instruments more than others. Further- more, it is assumed that no cost-sharing is required (as is 305 TABLE VII-5 Recommended NSF Program for Shared Instrumentation (Average Cost per Institution, $640K) Number of Universities Number of Instruments per Year Cost per Year ($) 20 (on average) 40 (on average) 60 (on average) 1 1/2 1/3 NSF FY 1985 departmental instrumental budget Needed increment 12.8M 12.8M 12.8M 38.4M 13.5M 24.9M

306 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES currently customary for dedicated instruments) and that support for mainte- nance and operation (infrastructure support) is assumed for a 3-year period. The outcome is an average NSF investment of $350K per instrument (see Table VII-6. The number of such instruments needed is reflected in the number of NSF grants made per year in those programs most dependent upon dedicated instruments. There is no point in making grants without providing the researcher his "tools of the trade." Four such programs are listed in Table VII-7 together with the planned FY 1985 budgetary allocation. To determine the instrumentation needs for international competitiveness, we assume that a reasonable fraction of the grants will be able to compete for one or two of these instruments every 6 years. We suggest that within a 6-year period, the top 10 percent of the grants might win two (different) instruments and that the next 20 percent would compete favor- ably for one instrument. As Table VII-7 shows, this im- plies that over a 6-year pe- riod, 268 instruments would be purchased, a rate of 45 instruments per year. This annual cost, $15.75M, ex- ceeds by more than a factor of two the amount recently planned. The planned FY 1985 NSF support of de~li- $220K cased instrumentation should be augmented by $9.6M. TABLE VII-6 International Competitiveness in Dedicated Instrumentation A dedicated instrumentation menu (1) Picosecond spectrometer (1) Molecular beam (2) FTIR (6) Lasers (Yag Doubled, Raman shifted) (1) High-resolution mass spectrometer (1) X-ray spectrometer (2) Dedicated computer (1) Array processor (15) Instruments Average capital cost per instrument Cost share, NSF, loon Maintenance and operation at 20% capital cost/year for 3 years Total cost, NSF share, per instrument 360K 250K 400K 540K 400K 500K 600K 250K $3300K 220K 130K 350K Grant Size As indicated earlier, instru- mentation has absorbed al- most all the growth in the Chemistry Division budget for the last 15 years. Though the instrumentation needs are by no means met, the impact on grant size has become increasingly acute. The adverse effect of this sustained level funding (corrected for inflation) is revealed by the level of effort supportable by the amount of an average grant. For example, if we begin with the proposed FY 1985 Chemistry Division budget, $92.IM, and subtract the amount planned for Departmental Instrumentation ($13.5M) and the probable amount for dedicated instrumentation (estimated at 10.5 percent, $9.7M), the remainder, $68.7M, is available for individual grants. The planned number of grants, 91S, implies an average grant size of $74.SK. Perhaps $30K will be extracted for institutional overhead and fringe benefits, which leaves only $45K for research personnel salary costs and research expenses. This amount will

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES barely support one postdoc- toral, one-and-one-half grad- uate students, and 1 or 2 months' summer salary for the principal investigator. Table VIl-S presents a range of possible grant sizes based upon various levels of effort and a set of reasonable assumptions about stipends, allotments for services, and supplies and small equipment needs. Overhead and fringe benefit costs are computed at an average value, 67 percent. The levels of effort A, B. and C proposed in Table VIl-S range from one postdoctoral student and six graduate stu- dents (level A) to one postdoc- toral and two graduate stu- dents ([eve] C). While a single grant at level A would not suffice for quite a number of our highly productive research groups, it defines a substantial research program. Level C, although a much smaller program, is still one with the vitality and continuity needed for effective fundamental research. Table VII-S adds levels D and E in recognition that research is also actively pursued and should be encouraged at undergraduate institutions and the smaller graduate institutions as an important element of undergraduate education and an effective means of invigorating faculty. 307 TABLE VII-7 Recommended NSF Program for Dedicated Instrumentation (Average Cost per Instrument, $350K) NSF Program Planned No. Grants FY 1985 Proposed FY 1985 Budget ($) Chemical physics Chemical dynamics Structural and thermochemistry Analytical TOTAL 180 180 180 130 670 For international competitiveness Top 10%, 2 instruments/six years Next 20%, 1 instrument/six years 268 Instruments Instruments _ AN ,.~ 6 years · _ ~1o Year 14.8M 17.0 15.5 11.0 58.3M 67.2 = 134 instruments 134.1 = 134 instruments 268 instruments $15.75M Present NSF budget = (0.105) ($58.3M) 6.12M (if held at 10.5% of program funds) Needed increment $9.6 M TABLE VII-8 Proposed NSF Chemistry Grant Sizes Level of effort A B C D E Postdoc, 20K 1 20.0 1 20.0 1 20.0 1/2 10.0 Grad's, 9K 6 54.0 4 36.0 2 18.0 1/2 4.5 Faculty summer, 2 mo 8.0 2 mo 8.0 2 mo 8.0 2 mo 8.0 2 mo 8.0 4K/mo Services, 3.5K (7) 24.5 (5) 17.5 (4) 14.0 (2) 7.0 (1) 3.5 Supplies, 4.5K (7) 31.5 (5) 22.5 (4) 18.0 (2) 9.0 (1) 4.5 Subtotal 138.0 104.0 78.0 38.5 16.0 Overhead fringe, 67% 92.5 69.7 52.3 25.8 10.7 Small instruments 8.0 6.0 4.0 2.0 2.0 Total 239K 180K 134K 66K 29K

308 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES With these levels in mind, we can now estimate the number of grants at each level needed to implement research programs that would maintain this country's international competitiveness, supply the talented scientific man- power sought by our industrial and government laboratories, and contribute our share to the increase of human knowledge. Table VII-9 shows a model with TABLE VII-9 NSF Grant Needs for International Competitiveness Number of Universities A B C D E Total 20 160 100 40 - — 300 40 60 120 160 — 340 60 10 20 30 60 120 Colleges ~ 80 80 230 240 230 60 80 840 Cost For international competitiveness 840 grants Present NSF Chemistry Division budget Less departmenal instrumentation Less dedicated instrumentation Present NSF chemistry division budget Less instrumentation Needed increment $55.0M $43.2M $30.8M $4.0M $2.3M $135.3M $135.3M $92.1M 13.5 6.1 72.5M $62.8M which to judge the present NSF grant program in the Chemistry Division. The model is consistent with and supported by the arguments that accompany Table VII-~. This mode] would provide NSF support for 2830 graduate students, about 35 percent of those currently engaged in Ph.D. study,- and 730 postdoctoral students, 26 percent of the number currently at this educational level. These percentages are reasonable when compared to the data of Table VII-2 showing that NSF provides 45 percent of the federal support for chemistry research performed in our universities and colleges. Of course, chemistry is supported elsewhere in NSF, which accounts for the remaining contribution NSF makes to the support of this essential science. Thus the proposed model points to a need for much larger funding in support of individual grants, even if the number of grants is not increased. That outcome does not imply or advocate focussing resources at a small number of established research centers at the expense of others. It does imply and advocate open competition for individual grants within the peer review system, and it assumes that an award, once made, will be adequate to complete the proposed research. It implies less proposal writing, less need for multiple grants to sustain a viable program, and provision for supporting infrastructure at all grant sizes. The discussion uses a tiered model merely to indicate the scope of the national

RESOURCES FOR BASIC RESEARCH lN THE CHEMICAL SCIENCES research effort that could and should be aggressively pursuing the opportunities presented in this report. It is not intended that NSF seek to establish such a tiered structure. Nevertheless, the tiered mode] dramatizes the fact that even with much larger funding for chemistry, the number of institutions that would be fully competitive on the international scene would be well within the aspirations and needs of this nation of 50 proud states. The startling revelation of Table VTI-9 is how far the proposed FY 1985 Chemistry Division budget falls short of these aspirations and needs. Table VIl-10 combines this with the analyses of Tables VIl-5 on shared instrumenta- tion and Table VIl-7 on dedi- cated instrumentation. The cumulative increment shows the magnitude of the discrep- ancy between needs for inter- national competitiveness and present support levels. Such inadequate support trans- lates directly into lost oppor- tunities and, in the long run, into lost leadership in science and industry. The nation can- not afford these losses. 309 TABLE VII-10 NSF Chemistry Division Budget Incremental Needs for International Competitiveness in Today's Frontiers of Chemistry FY 1985 Increment Request Needed Total Need (in FY 1985 dollars) $38.4M 15.7 135.3 Departmental instrumentation Dedicated instrumentation Grants Total $13.5M 6.1 72.5 $92. 1M $24.9M 9.6 62.8 $97.3 $189.4M Table VIl-10 shows that the NSF Chemistry Division budget should be doubled, after correction for inflation. For cost-effectiveness, it is recommended that the NSF begin a 3-year initiative to raise the NSF support for chemistry by 25 percent per year for JAY 1987, FY 1988, and FY1989. The added increment should be used for increasing grant size, ensuring encouragement of young investigators, enhancing the departmental and regional instrumentation pro- gram, and increasing the amount for dedicated instrumentation. Creativity Within NSF The Chemistry Division in NSF has a laudable record of innovativeness and flexibility in its funding patterns. Its Young Investigator and Shared Instru- mentation programs were responsive to encouragement by the 1965 predecessor to this report. More recently, it has experimented with new contractural mechanisms that cluster grants from a given institution, peer review based on "recent track record," one-time extensions based on creativity, and Regional Instrumentation Facilities. The Chemistry Division is encouraged to continue seeking new mechanisms of funding that help NSF fulfill its unique role of stimulating adventurous and innovative research with as little constraint on that research as possible. In the simplest analysis, NSF should try to support creative people in a fashion that permits them to be creative. This can be done within the peer review system by eliciting judgments from reviewers based as much on their evaluations of the investigator's research promise as on his or her research

310 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES promises, as stated in the proposal. For established investigators, considerable weight should be placed on originality evident in recent accomplishments. On this basis, perhaps 10 percent of the existing grants might be extended a few years without a new proposal (at the program officer's discretion). (This is now being done in NTH with the so-called "Javitz grants" within the Institute for the Neurosciences). Alternatively, especially creative programs ready for renewal might be supported for longer durations, such as ~ to 7 years. The NSF should maintain its attention to the special needs of the new investigators whose continued entry into chemistry is essential to its long-range vitality. A current difficulty is the large "start-up" costs associated with initiation of research in almost any area of chemistry. A second problem is the time required for build-up of research productivity. Both require risk-taking that must be open to NSF program officers. Grant duration must be sufficient to avoid forcing young scientists into pedestrian research programs sure to lead to quick publications; 4- and 6-year grants should not be considered excessive. Multiyear grants should accommodate especially large, first-year "start-up" costs. Finally, within bounds set by reasonable accountability, NSF should strive to minimize the number and length of proposals required for an active and productive investigator to obtain adequate support. The increase in grant size advocated here would immediately reduce both the number of proposals per investigator and the diversion of the research community in the operation of the peer review system. Increased grant duration would have the same salutary effects. Grants of less than 3-year's duration should be avoided, and 5-year grants should not be uncommon. Finally, the current NSF limit on proposal length about 15 single-spaced pages for the proposal itself should be en- forced. Both program officers and reviewers should be encouraged to return single-investigator proposals that are excessive in length. CHEMISTRY AND THE DEPARTMENT OF ENERGY MISSION DOE support of the chemical sciences is embedded in a total R&D FY 1985 budget amounting to $577SM, which rose by 4.] percent from the FY 1984 level. However, 44 percent of this is directed toward Defense R&D, which rose by 10.3 percent. When these R&D activities are subtracted, the non-Defense R&D total, $3245.9M, is found to fall slightly relative to FY 1984 (by .2 percent). In another profile of the FY 1985 DOE R&D budget, two-thirds will be expended in the DOE National Laboratories, which, all together, saw their proposed collective budgets rise by 8.5 percent from 1984. By difference, the one-third expended outside the National Laboratories ($1889.SM) fell by 3.S percent. A component of the difference is attributable to the concentration of the overall budget increase in Defense-related R&D, most of which is conducted in 4 or 5 of the 12 National Laboratories.

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES Need for a DOE Initiative in Chemistry With these 1984-1985 changes in mind, it is appropriate to assess the relative emphasis placed on fundamental research in chemistry by DOE in pursuit of its mission. To do so, we have listed in Table VIl-ll the FY 1985 budgets of the most fundamental research programs conducted under the DOE headings "Energy Supply" and"General Science and Research." The individ- ual programs are listed in or- der of size, and the last col- umn shows the percentage of the total for those programs. The table shows that of the programs listed, the Materi- als Sciences represent 10.9 percent, the Chemical Sci- ences 5.3 percent, and the Nu- clear Sciences (where Nuclear Chemistry is found) 2.5 per- cent. The Materials Sciences Pro- gram devotes a modest per- centage of its budget to chem- . . . . ~ TAB LE V I I- 1 1 Magnitude of Sel ected D OF Fundamental Research Programs, FY 1985a High energy physicsb Magnetic fusion energy Materials sciencesC Environmental sciences Nuclear physicsb Chemical sciencesC Nuclear sciencesC Applied mathematical sciencesC Engineering and geosciencesC Biological energy researchC Advanced energy projectsC Total $547.8M 440.1 193.6 191.1 180.6 95.1 44.0 36.8 28.3 13.1 11.1 $1781.6M 30.7~o 24.7 10.9 10.7 10.1 5.3 2.5 2.1 1.6 0.7 0.6 aAAAS Report IX: Research and Development, FY 1985 (following congressional action, as compiled by AAAS in Nov. 1984). b Programs conducted under "General Science and Research." c Programs conducted under "Basic Energy Science" within "Energy Supply." lstry-orlented programs (about 7 percent) including programs on Polymers (1.3 percent), Catalysis (1.9 percent), Corrosion (3.0 percent), and Combustion (.4 percent). About 9.4 percent of the Nuclear Science budget goes to support heavy-element chemistry. One can explain the sense of urgency implicit in the very large annual investment being made in the development of nuclear fusion, an investment that has been sustained now for almost two decades. It is quite impossible, however, to rationalize the extremely modest emphasis placed on chemistry, as embodied in the 5.3 percent figure. Chemistry and Chemical Engineering will be the dominant sciences in the energy technologies that must be developed and that we will depend upon for the rest of this century. In 1983, 92 percent of the U.S. energy consumption was based upon chemical fuels. The small remainder, ~ percent of our use, was mostly provided by about equal contributions from nuclear and hydroelectric energy. As the DOE contemplates its mission, to assure future access to abundant and clean sources of energy, it must reckon with these challenging expectations for the next three decades: 311

312 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES by the year 2000 the U.S. annual energy consumption will probably exceed today's use by 20 to 50 percent; Curing the next three decades, growth in the use of nuclear power will be severely constrained by social concerns already in evidence; further increase in hydroelectric power has natural limits and is in conflict with widespread desire to minimize environmental encroachment; even the most optimistic proponents of nuclear fusion do not see it providing a large fraction of our energy use before well into the 21st century; Dependence upon high-grade petroleum crudes and high-grade coal deposits must decline as worldwide reserves are depleted and as access to foreign crude oil is capriciously restricted by political developments beyond our control. These boundary conditions surely point to the need for expanding the knowledge base upon which new energy technologies can be built. Chemical and electrochemical systems provide some of the most compact and efficient means of energy storage. And we can predict with confidence that foremost among the new energy sources will be low-grade chemical fuels, such as high-sulfur- content coal, shale oil, tar sands, peat, lignite, and biomass. For not one of these alternatives does appropriate technology exist today that can economically meet the stringent demand that environmental pollution be avoided despite the higher impurity content of these new feedstocks. Enormous chemical challenges must be met—for new catalysts, new processes, new fuels, new extraction techniques, more efficient combustion conditions, better emission controls, more sensitive environmental monitoring, and many others. Biomass must be brought to practicality to reduce the amount of fossil fuel burned and thus to help check the rate of increase in atmospheric carbon dioxide. (Combustion of biomass grown for fuel cannot return more CO2 to the atmosphere than was consumed in the photosynthetic processes by which the biomass was produced). Solar energy must be fully exploited including through development of artificial photosynthetic and electrocatalytic techniques that avoid combustion by con- verting the photon energy directly to electrical or chemical energy. To acceler- ate our movement toward meeting these critical needs the Department of Energy should mount a major initiative in those areas of chemistry relevant to the energy technologies of the future. The same urgency that justifies our powerful program in nuclear fusion dictates a program of comparable magni- tude in chemistry. The scope of the initiative should be broad, encompassing all forms of catalysis (homogeneous, heterogeneous, electro-, photo-, and enzymat- ic-catalysis), exploiting the new potentialities for understanding combustion, developing the plant sciences that will facilitate our use of natural photosyn- thesis, and adventuring boldly into frontiers that may open new energy avenues (genetic engineering of photosynthetic organisms, development of polymeric photovoTtaic devices, etc.J. The program should take full advantage of the multidisciplinary capabilities of our National Laboratories. Although its fund- ing is set at a modest level, the Combustion Research Facility at Sandia

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES Livermore provides a laudable example. In an appropriate number of our National Laboratories, the defined mission should be reshaped to inclucle a major focus on one or more of the areas of chemistry crucial to energy technologies. And, most important, the DOE research support to universities must be increased to engage the larger chemistry community in the accomplish- ment of the Department's mission. While the chemistry activity at the National Laboratories is being increased, university research programs in energy- relevant areas of chemistry should be raised to be commensurate with those in the National Laboratories. In addition, the interaction between DOE laborato- ries and universities should be strengthener' through visiting faculty re- searcher programs and long-term collaborative projects. A Proposed 5-Year Initiative A program to approach these goals in a cost-effective way is presented in Table VIl-12. It proposes to build toward a level of commitment about half as big TABLE VII-12 A Proposed 5-Year DOE Initiative in Chemistry (All in FY 1985 Dollars) Biological Energy + Chemical Sciences + Incremental National Catalysis + Combustion Growth Labs Universities FY 1985 $120M $ $ 94M $ 26M 1986 142M +22M lOlM 41M 1987 169M + 27M 110M 59M 1988 201M +32M 121M 80M 1989 239M + 38M 134M 105M 1990 285M + 46M 149M 136M Total growth over 5 years $165M as that for magnetic fusion, appropriate to the reality that 90 percent of our energy must come from chemical energy sources for the next quarter-century. Incremental growth by a factor of about 2.5 is needed! in chemistry research programs relevant to the DOE mission. This level of effort should be roughly equally divided between the National Laboratories and university laboratories. The substantial growth implied is approached over a 5-year period at a real growth rate of 20 percent per year for 5 years to ensure cost-effective distribu- tion of the incremental resources and to provide orderly growth of the DOE program management expertise that will be needed. Each yearly increment in this 5-year program would be divided in a growth pattern that will, first, stimulate the efforts of those National Laboratories that choose to refocus their defined mission toward chemistry-based energy technologies and, second, build the engagement of the academic research community in the areas that undergird these technologies. The FY 1985 baseline is taken to be $120M, based upon the approximate present sum of the Chemical Sciences, the Biological 313

314 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES Energy Research, and the Catalysis, Corrosion, and Combustion programs. This initiative is intended to be broad in scope to encompass the range of chemistry frontiers at which advances can be expected and which are obviously of relevance to the development of new energy technologies. A desirable distribution of the incremental resources is given in Table VII-13. TABLE VII-13 A Desirable Deployment of Incremental Resources (All in FY 1985 Dollars) 5-Year Program 1986 1987 1988 1989 1990 Growth Catalysis Heterogeneous $ 6M $+6M $+7M $+8M $+ lOM $+37M Homogeneous +2M +3M +4M +5M +6M +20M Enzymes and artificial +2M +3M +4M +5M +6M +20M enzymes Electrocatalysis +3M +4M +4M +5M +6M +22M Reaction dynamics +3M +4M +5M +6M +7M +25M Photosynthesis and +3M +4M +4M +5M +6M +22M photochemistry Plant sciences +3M +3M +4M +4M +5M +19M Increment $22M $27M $32M $38M $46M $165M Technica] Qualifications at the DOE Nationa] Laboratories In light of our recommendation that the DOE mount an initiative in support of chemistry-based energy technologies, it is useful to ask about the distribution of technical qualifications of the scientific staffs at the existing National Laboratories. Furthermore, changes over the last decade in professional quali- fications at each of the National I~aboratories may indicate changing trends in the way each Laboratory sees its mission evolving. Table VII-14 shows for four National I~aboratories the number of physicists and of chemists in 1970, 1980, and 1984. In assessing these figures, it is perhaps relevant to remember that U.S. business employs approximately 2.5 times more Ph.D. chemists than Ph.D. physicists. The ratios shown in the last column of Table VII-14 are revealing. The only trend discernible is at Oak Ridge where the hiring pattern must reflect a deliberate movement away from chemistry-based technologies. At Argonne and Brookhaven, the hiring pattern seems to signal a general laboratory expansion over the 1970s followed by a retreat or slowdown in growth. However, in neither laboratory is there evidence of significant change toward chemistry-based mission goals. For Los Alamos only current data were available, and they display heavy emphasis on physics-oriented goals. These compositions imply leanings away from chemistry-based technologies, an outcome more or less dictated by and appropriate to the distribution of DOE research resources presented in Table VII-ll. The program advocated in Table VII-12 projects an

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES TABLE VII-14 Distribution of Chemists and Physicists at the DOE National Laboratories National Number of Ph.D. Professionals Number of Physicists Laboratory Year Physicists Chemists Number of Chemists Argonne 1971 168 128 1.31 1980 191 +23 159 +31 1.20 1984 169 132 1.28 Oak Ridge 1970 178 +72 180 _4 .99 1980 250 + 19 176 - 10 1.42 1983 269 166 1.62 Brookhaven 1970 200 +60 82 +51 2.50 1980 260 133 1.95 1984 276 + 16 134 + 1 2.06 Los Alamos 1984 679 242 2.81 overall growth by about 65 percent of the chemistry-oriented research activity at the National Laboratories. It is recommended that growth in chemistry- oriented research should be concentrated at those National Laboratories that choose to refocus their defined mission toward chemistry-based energy technol- ogies. This growth should add substantially to the cadre of Ph.D. staff with credentials in areas of chemistry appropriate to the new mission. CHEMISTRY AND THE NIH MISSION According to AAAS Report ~X: Research and Development, FY 1985 (see Table Il-15, p. 132), the NIH budget in FY 1984 was $4477M, more than 3.5 times the NSF budget for the same year. Of this, $637M was directed to NIH intramural research and $2387M to extramural research projects. Thus, these two elements together account for, respectively, 12.0 and 53.3 percent of the total budget. In contrast to other federal agencies that support chemical research (NSF, DOE, and DOD), not one of the National Institutes of Health has a program expressly titled "Chemistry." Neither are there any program descriptions that identify as a primary program goal the advance of particular subbranches of chemistry that could be seen to be particularly relevant to the Institute mission. Nevertheless, NTH furnishes a substantial fraction of the U.S. federal support for university research in chemistry, surely in excess of 25 percent. Thus, NTH plays a crucial role in support of basic chemical research in this country. The fact that chemistry is not overtly supported makes it difficult to track the magnitude and distribution of the substantial NIH funding of research con- ducted by chemists. This is evident in the 1983 NSF study of federal research support to chemistry, agency by agency (see Appendix Table A-7. The table displays FY 1982 amounts directed toward basic research in chemistry con- ducted at the nation's universities: by NSF, $76.4M, and by NTH, $38.3M. The 315

316 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES NSF figure can be derived as a fraction of the announced budgets for its divisions; the sum of the NSF Chemistry Division budget and the chemistry part of the Materials Science Division budget is $77.3M (see Table VIl-3), and smaller amounts are derived from the Divisions of Biochemistry, Atmospheric Chemistry, Geochemistry, and Marine Chemistry. Contrast this agreement with the outcome of a more detailed NTH budget analysis sponsored by the Board on Chemical Sciences and Technology of the National Research Council for FY 1982. This analysis was not based on the stated aim of the research (which, perforce, must always be justified in terms of human health) but rather on the total funding of individual investigators whose institutional connection is a chemistry department at a Ph.D.-granting univer- sity. The total so obtained was $76M, just double that recorded in Appendix Table A-7. The discrepancy undoubtedly reflects the substantial cross- disciplinary character of chemistry research as conducted in our university chemistry departments and implicitly justifies the substantial support chemis- try receives from NTH. Average Grant Size Table VIl-15 shows some detail on the FY 1982 distribution among the institutes of the funding received by chemists. About 90 percent of the grants are supported by ~ of the 11 Institutes and 54 percent by the Institute for General Medical Sciences. Over all the institutes, the average grant size is $83.2K. This is only slightly larger than the NSF average, $74.SK, and it again will support less than an average level of effort C in Table VIl-S (one postdoctoral and two graduate students). By the same argument made earlier concerning NSF grant sizes, it is recommended that a fraction of any additional NIH funds into chemistry be used to increase average grant size. A 30 percent increase in average grant size to level of effort B (Table VIT-~) is a reasonable target. Furthermore, somewhat larger grant size should be considered appro- priate for cross-disciplinary collaborative programs that, through joint PI structure, link expertise in chemistry with that in other disciplines (biology, molecular biology, etc.~. Grant Success Ratio Through all of its institutes in FY 1982, NIH supported 14,826 individual investigator grants with a total cost of $~.45B. Thus chemistry receives about 6 percent of the research monies distributed by NTH through individual grants and about 3.7 percent of the total resources NTH directs to the conduct of R&D at universities and colleges (in FY 1982, $2.07B). When the data base is expanded to include, as well, departments of biochemistry, pharmacology, and medicinal chemistry, total NTH support in the form of grants to individual investigators rises to some $204M in FY 1982 with 2,256 individual grants. To these four departments, the Institute for General Medical Sciences devotes nearly 40 percent of its budget, over $94M, to grant support. Support from the

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES National Cancer Institute rises for these four departments to $2SM, somewhat less than 10 percent of the total NCI research budget. The last column of Table VIl-15 shows that about 6 percent of all NTH individual investigator grants are made to investigators in university chemis- try departments but that the percentage is considerably higher in the Institute for General Medical Sciences, 19.3 percent. For this particular Institute, detailed study of awards made relative to the number of applications received over the period 1974-1982 shows that the percentage of applications awarded has consistently remained a few percentage points above the rate to nonchemistry scientists over this period. Thus scientists in chemistry depart- ments are apparently being equitably treated as measured by the success of their applications. On the other hand, the average success rate for competing research projects has declined from about 60 percent in 1974 to about 40 percent in 1982. Equally damaging is the fact that success rates fluctuate widely from year to year, moving from about 60 percent in 1975 to 40 percent in 1977, back up to 55 percent in 1979, and then down to 37.5 percent in 1982. Such large and apparently capricious variations do not afford the continuity essential for first-cIass fundamental research. Needless to say, the overall NTH budget did TABLE VII-15 National Institutes of Health Research Projects to Chemistry Department-Based Principal Investigators, FY 1982a Percent of Total Approximate Number, Number, Res. Institute Dollarsb Research Grants Grants General Medical Sciences 42.9M 499 19.3 National Cancer Institute 11.7M 152 5.6 Allergy and Infectious Diseases 4.5M 55 3.8 Heart, Lung, and Blood 6.2M 67 2.9 Arthritis, Diabetes, Digestive, 5.3M 64 2.7 and Kidney Diseases Neurological and Communica- 1.8M 25 1.7 five Disorders and Stroke Environmental Health Sciences 1.2M 18 5.5 Aging .4M 5 1.2 Dental Research .4M 5 1.4 Eye Institute 1.7M 23 2.3 Child Health and Human De- .7M 10 0.8 velopment Total all institutes 76.8M 923 5.6 a Robert M. Simon, National Research Council, private communication. b These dollar amounts do include institutional overhead. not display swings of these magnitudes, so other factors are at work in producing these large amplitude fluctuations. Whatever the causes, these rapid 317

318 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES NATIONAL INSTITUTES OF HEALTH 602 At: At: 40% 20% //:INDIVIDUAL INVESTIGATOR/// _ /~/j//-RESEARCH GRANTS///////, /~i//~.~i: ~ ~ _ 1 975 1 980 YEAR NIH AWARD RATES ARE DECLINING AND FLU CTU ATING WIDELY changes disrupt and damage high-quality research pro- grams, and they are to be avoided. Hence we applaud anti vigorously support the ef- forts of the National Insti- tutes of Health to build into its yearly budget an extramu- raZ grant stabilization pro- gram. , ~ , Shared Instrumentation 985 In the 1960s, NIH began a program of support for the purchase of large instrumen- tation—mainly NMR and Mass Spectrometers- for shared use at universities. The program stabilized in the support of existing centers without new starts in the early 1970s and gradually was phased out. Nevertheless, NIH had furnished an admirable prototype mode] for the NSF Departmental Instrumentation program which, from a much smaller funding base, attempted by itself to meet chemistry departmental needs through the 1970s. Now NIH has reawakened its shared instrumentation program, a timely addition to the existing oversolicited counterpart programs in NSF and DOD. In view of the growing dependence upon sophisticated instrumentation in the health-related sciences, we recommend that NIH maintain its extramural shared instrumentation program at a [eve! approximately equal to that pro- posed here for NSF. Further, we urge that initial cost-sharir~g and support of ongoing maintenance and operations follow the same guidelines proposed earlier for NSF, i.e., 80 percent cost sharing and 5-year maintenance at 20 percent of the initial cost. CHEMISTRY AND THE DEPARTMENT OF DEFENSE MISSION The many ways in which our national security depends upon a vigorous scientific enterprise are enumerated in the introduction in Section V-C. The body of this report presents compelling evidence for the role that basic and applied chemical research has played in this enterprise, with direct impact on the technologies upon which the nation's security is based. Most obvious examples of interest to DOD include fuels and propellants, new polymers for protective garments, structural elements, and nose cones, alloys for jet engine parts, electronic materials, upper atmospheric chemistry, medicines, chemical lasers, and a host of others. Plainly, it is appropriate for the Department of

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES Defense to devote a deliberate fraction of its resources and attention to the maintenance of the research activity that leads to such advances. Applied and Basic Research For near-term exploitation of scientific advances, DOD should invest heavily in applied research, which is designated by categories 6.2, 6.3, and 6.4. At the same time, long-range security interests dictate that DOD should also support fundamental research in those areas that can be seen to underlie technologies of particular relevance to the defense mission. The more fundamental research is designated category 6.1. Over the last two decades, there have been profound changes in the pattern of DOD research activity. When corrected for inflation, the level of support for all R&D declined steadily from 1970 to 1975, remained constant until 1980, and then was raised steeply during the last 5 years. In 1984 the level of activity in applied research exceeded the 1965 level and in the proposed 1985 budget, it is 22 percent higher. When attention is focussed on basic research (6.1), the picture is remarkably different. Viewed as a percentage of the total R&D, in 1965, 5.1 percent of the R&D was directed to fundamental research. In 1980, this figure had dropped to 3.9 percent. In 1984, the 6.1 category accounted for only 3.0 percent of the total R&D, and in the proposed 1985 budget, it drops still further, to 2.7 percent. Expressed in level of effort in inflation-corrected dollars, (CORRECTED FOR INFLATION TO 1972 S) the proposed 1985 investment ~ i` in basic research is only two- 0 So 0 B ~ <~_ r 1 Offs r I thirds the 1965 value. It is I, recommended that over the ° next 5 years, the percentage <: $5B of the DOD R&D budget dEi- ° rected to basic (6.1J research be increased to restore the 1965 value of 5.0 percent by the year 1990. Research Areas A number of broad research areas in chemistry deserve DOD attention because they are likely to provide signifi- cant advances relevant to de- fense technologies. Strategic and Critical Materials —Fuels, Propellants, and Explosives S600M an: ~ 5400M son ct ~ $200M 965 ~ 970 ~ (375 1 980 ~ 985 YEAR D.O.D. SUPPORT FOR APPLIED RESEARCH (~.2 6.3 6.4 (1985 LEVEL)/(1965 LEVEL) = /5 (CORRECTED FOR INFLATION TO 1972 S) _ ( 196 5 LEVEL ) i, _~ 1 965 1 970 1 975 1 980 1 985 YEAR D.O.D. SUPPORT FOR BASIC RESEARCH (6.1 ) ( l 98 5 LEV EL )/( 1 9 65 LEY EL ~ = /3 319

320 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES Atmospheric Phenomena —Chemical and Biological Defense Nuclear Power and Nuclear Weapons Ejects For each of these broad areas, Table VIl-16 lists fundamental studies that will TABLE VII-16 Defense Needs and Special Opportunities Army Navy Air Force DARPA Strategic and critical materials Polymers as structural materials Solid state chemistry Chemical synthesis Fuels, propellants, and explosives Molecular spectroscopy & kinetics Chemical synthesis Catalysis, surface sciences Combustion Corrosion Chemical lasers Fluid transport Condensed phases Theoretical chemistry Atmospheric phenomena Chemical kinetics Atomic & molecular spectroscopy Analytical chemistry Theoretical chemistry Chemical and biological defense Biotechnology Analytical chemistry Marine chemistry Organic synthesis Nuclear power and nuclear weapons effects Nuclear chemistry and nuclear processing Nuclear stability Avow '' Eva ~v~v~ vet vet v. Eva '' vet v. vet v. ~v. '' vet ~ vat vet ~ vial NOTE: ~ = applicable '' = important Nave = critically important provide the advances that will lead, in the long run, to advanced defense concepts and applications. The table also indicates specific applicabilities to the specialized interests of the various defense arms. Chemistry Research in Universities With so many areas of opportunity of special relevance to the DOD mission, there is ample reason for DOD to guarantee the vitality of U.S. research activity

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES in chemistry. Furthermore, there are clear benefits to DOD if these opportuni- ties are pursued, in part, in our university research laboratories. First, univer- sity participation builds the technical manpower pool needed to deploy and maintain our increasingly sophisticated defense technologies. Second, it gives DOD significant influence on the university research agenda in directing attention toward advancing our knowledge in those areas of chemistry key to our defense posture. Plainly, the extent of DOD influence on university research agendas is related to the fraction of the university support coming from DOD sources. This fraction remained close to 10 percent through the decade of the 1970s, a level much too Tow to secure lithe desired end. In order for DOD to have a significant impact on building our technical manpower pool while increasing the growth of critical scientific knowledge, its support of chemistry must be comparable to that of other federal agencies. Since four agencies furnish most of the basic research support for the chemical sciences, a reason- able level for DOD support is near 25 percent of the total. In fact, beginning in 1981, the percent of DOD support has been growing. From 1980 to 1983, DOD funds for uni- versity research in chemistry rose from $15.0M to $29.5M, sufficient to bring this percent to 16 percent. This rise paral- ~ ~~ ~ ~ · ~ 321 (corrected for inflation to 1972 S's) 10~: 25~ Federal Support for University Research in Chemistry Percent from Dept. of Defense a' .. .. ... .. ~ , . . 1. 1 1 1 1975 1 980 1 985 YEA R TECHN ICAL MANPOWER AND CRITI CAL KNOWLEDGE REQUIRE A LARGER D.O.D. INVESTMENT Select a corresponding much-needed rise in in-house research investment from $20.7M to $30.4M, which, after inflation, represents a real growth of 4 percent per year. Then, in the years 1984 and 1985, planned support for chemistry research leveled oh, just matching inflation at 6.1 percent per year. It is recommended that DOD support for university research in the chemical sciences should be raised to about 25 percent of the total federal support. Real growth of 10 percent per year should be sustained until that goal is reached. Because of the central importance of chemistry to our national security, the same proposed growth should be Provided to DOD in-hou.se re.search nroarr~m.s of the 6.1 category. 6.1 Research in the Various Defense Arms ~ ,~ Table VIl-17 shows for the years 1982 to 1985 the total DOD investment in fundamental research, the percentage of these amounts devoted to chemistry and materials programs, and how these amounts were divided among the

322 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES various arms. The Army, Navy, and Air Force direct approximately equal amounts to the support of chemistry, which is appropriate in view of the opportunities chemistry affords (see Table VIl-161. In the preceding discussion, it was recommended that the benefits to DOD warrant an increase in its current level of support for chemistry of 10 percent real growth per year for 5 years. For each of the three defense arms, this growth should be focused into the high pay-offresearch areas identified in Table VII-16. Comparable growth in the Chemistry and the Materials Science programs should exploit the opportunities chemistry offers to provide new strategic materials, fuels, propellants, and explosives, as well as deeper understandings of chemistry relevant to atmospheric phenomena, biological defense, and nuclear power and weapons elects. Collaborative Relationships Since it is to the benefit of both the inhouse DOD laboratories and the broader chemical research community (including both industry and universities), it is recommended that means be sought to increase the interaction between DOD laboratories and universities. There are several mechanisms that should be pursued to improve these interactions; they include (a) postdoctoral and visiting faculty programs, (b) long-term collaborative projects, and (c) innovative grad- uate student support programs. Such interactions are entirely consistent with recommendations of both the Grace Commission and the Packard Committee bearing on the strengthening of federal laboratory, university, and industry interactions. TABLE VII-17 Department of Defense Research Support for ChemistrY and Materials Science, 1982-1985 (6.1, Basic Researches 1982 1983 1984 1985 (Request) DOD Total $694.1M $780.0M $837.9M $897.9M Chemistry 53.1 (7.7%) 58.9 (7.6C%o) 62.0 (7.4%) 66.3 (7.4%) Materials 71.5 (10.3%) 81.0 (10.4%) 82.8 (9.9%) 87.5 (9.7%) ARMY Total $178.1M $202.4M $217.5M $238.8M Chemistry 22.2 (12.5%) 23.8 (11.6%) 23.5 (10.8%) 25.4 (10.6~o) Materials 14.8 (8.3%) 17.1 (8.4%) 17.1 (7.9%) 18.1 (7.6%) NAVY Total $276.0M $309.3M $319.3M $349.7M Chemistry 17.1 (6.2%) 18.2 (5.9%) 19.0 (6.0%) 20.9 (6.0%) Materials 24.2 (8.8%) 24.0 (7.8~o) 25.0 (7.8%) 27.5 (7.9%) AF Total $147.5M $167.3M $192.5M $206.9M Chemistry 13.8 (9.4%) 16.9 (10.1%) 19.5 (10.1%) 20.0 (9.7%) Materials 18.1 (12.3%) 17.7 (10.6%) 21.9 (11.4%) 22.7 (11.0C/o) DARPA Total $ 92.5M $101.0M $108.6M $102.5M Materials 14.4 (15.6%) 22.3 (22.1%) 18.8 (17.3%) 19.2 (18.7%) a AAAS Report IX: Research and Development, FY 1985.

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES Equipment and Facilities Support It is now widely recognized- as documented at length in this report that federal programs supporting science and engineering activities have not ade- quately recognized the sophistication and cost of the equipment required for modern science. In many cases the equipment found in U.S. laboratories supported by DOD 6.1 and 6.2 programs is greatly outdated compared to the equipment found in laboratories undertaking similar work in U.S. industry or in foreign countries. In 1983 DOD recognized these equipment needs by establishing a special "set aside" program for instrumentation, $30M per year divided equally among its three arms. This heavily overso]Licited program has proven to be quite successful, except that maintenance and technician support have not been included as an integral part of the equipment award. Since recipient research groups and their institutions do not have adequate resources to devote to these important purposes, there is a risk that the much needed new instrumentation will be less efficiently employed or will become obsolete more rapidly than would be the case if provision were made to include appropriate levels of maintenance and technician support. It is recommended that DOD continue its instrumentation program but with the addition of support for maintenance and operation to ensure cost-effective use of the equipment. This instrumentation program should not grow at the expense of the direct contract support for research activities. Evidently the appropriate balance between research support and equipment support is a matter requiring on-going judg- ment, and continuing attention must be given by DOD to maintaining this balance. With regard to research facilities at universities and colleges, it is regrettable that the lapse in federal programs has meant that new research laboratories are not being built and that old laboratories and buildings are deteriorating progressively. While the general state of research buildings and laboratories on U.S. campuses is clearly not a principal responsibility of DOD, the Department does have an interest in assuring that campus research facilities are adequate to carry out the research missions and associated technical manpower training for long-run national defense. Accordingly, DOD should explore mechanisms to support new construction and renovation of university research facilities in particularly critical areas of chemical science. If attention to such space needs is not forthcoming, it cannot be expected that an adequate research base for DOD needs will be available. In addition, DOD should support OSTP efforts to establish new programs for research facilities generally at the nation's colleges and universities. CHEMISTRY AND THE DEPARTMENT OF AGRICULTURE MISSION It is widely recognized abroad that a major strength of the U.S. fundamental research enterprise is the plurality of federal agencies sponsoring it. Each 323

324 RESOURCES FOR BASIC RESEARCH Ix THE CHEMICAL SCIENCES agency exerts its due influence on the research community's agenda by encouraging research at the most fundamental level in areas that underlie and advance that agency's mission. Consequently, it is an unfortunate anomaly and a loss to the nation that the Department of Agriculture has had difficulty mounting a significant competitive grants program to engage the university research community more fully in its mission goals. The FY 1984 R&D budget for the USDA was $869M, 1.9 percent above the FY 1983 figure. Of this, only 4.5 percent is directed toward research in chemistry ($38.7M in FY 19831. The bulk of this research is performed in-house in seven major research centers under the Agricultural Research Service. In fact, as shown in Table A-7, less than one-fifth of USDA chemistry research is supported through competitive grants at universities and colleges ($6.6M in FY 19831. The result is that USDA provides only about 3.6 percent of the federal support for chemistry performed in the nation's university research laboratories. This is incongruous in the light of chemistry's significant accomplishments relevant to increase in the world's food supply (e.g., in fertilizers, growth hormones, pesticides, herbicides, pheromones, and genetic engineering of plants) and the expanding possibilities for further advances as described in Sections IV-A and IV-D. Over the last several years, conscientious and laudable attempts have been made by USDA to add to its budget a more substantial competitive grants program that does not detract from the existing activities of the Agricultural Research Service. These attempts have not yet been implemented by Congress. In the public interest, it is recommended that the Department of Agriculture initiate a substantial competitive grants program in chemistry research. The aim of the program should be to increase over the next 5 years the Department's extramural support of fundamental research in chemistry to an approximate par with its intramural research program. CHEMISTRY AND THE NATIONAL AERONAUTICS AND SPACE ADMINISTRATION MISSION This report identifies several rapidly moving fronts of chemistry in scientific areas where NASA has active interests. Obvious examples are high energy propellants (Initiative A), high temperature materials (Initiative E), chemistry in plasmas (Initiative E), and chemistry in the stratosphere (Initiatives A and D). The current NASA interest in the establishment of a permanent human presence in space offers fascinating challenges in chemistry related to life- sustenance within the closed system of a space station (Initiative C). It is recommended that NASA direct increased attention toward special opportuni- ties in chemistry relevant to operations in space: high-energy propellants chemical behavior under extreme conditions

RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES reaction kinetics and photochemistry under collision-free conditions chemical aspects of life-sustenance in a closed system analytical methods for compositional monitoring in both the troposphere and the stratospere Especially vital is NASA's concern with the chemical composition of the atmosphere, particularly in reference to changes that may be occurring. The importance of a deep understanding of the complex chemical processes operative in the stratosphere has already been well documented in recent concerns about the ozone layer. Now attention is focussing on the atmospheric carbon dioxide concentration, which has clearly increased over the last three decades. Similar changes are being noted in methane and nitrous oxide concentrations, each of which is intimately involved with the earth's biological activity. Understanding of the biogeochemical cycles that move carbon, nitrogen, sulfur, and other elements into and out of the atmosphere is fundamental to the maintenance of favorable conditions for all forms of life. Because of NASA's unique capabilities, it can play a significant role as we seek this understanding. NASA has been actively engaged in such studies, through downward-Iooking sensors in satellites, through programs of active measurement in the atmo- sphere, and through laboratory-backup programs in chemical kinetics. These programs address such crucial problems as the role of trace gases in the atmospheric trapping of infrared radiation (the greenhouse effect), the deposi- tion and chemical fate of acid compounds and their precursors, the chemical interactions that affect stratospheric ozone, and the environmental implications of the changing desert cover in the tropics. Clearly, NASA should maintain a substantial and continuing commitment to the study of atmospheric chemistry. NASA conducts a large and productive research program through its own NASA laboratories, and it depends significantly upon private research contrac- tors. It engages less fully the academic research community in chemistry. In light of the potential contributions of chemistry to the safety, range, and effectiveness of future space operations, NASA should more actively encourage academic chemists to address problems relevant to the NASA mission through competitive grants for funclamental research. CHEMISTRY IN THE ENVIRONMENTAL PROTECTION AGENCY For FY 1985, EPA proposed to direct 6.5 percent of its $4.25 billion budget request toward R&D. Approximately half of this $27SM R&D support would go to programs in which chemistry plays a role: Environmental Engineering and Technology, $50M; Acid Rain, $34M; Monitoring Systems and Quality, $31M; Exploratory Research, $16M; and Environmental Processes and Effects, $1.9M. While the Exploratory Research program is only 5.6 percent of the total EPA R&D program, over 95 percent is currently directed toward extramural activity. This research is largely conducted at eight university-based centers, but it also 325

326 RESOURCES FOR BASIC RESEARCH IN THE CHEMICAL SCIENCES includes some competitive research grants. Significantly, this small program was incrementally increased by a factor of 2.5 in FY 1984 over the FY 1983 level, and all the growth was placed in the extramural program. Presuming that this Exploratory Research program is intended to nurture long-range research relevant to its mission, EPA should increase the percentage of its R&D funds placed in its Exploratory Research program and its commitment to extramural fundamental research relevant to environmental problems of the future. Most of this growth should be awarded through competitive grants. Initiatives A and D both present opportunities that are applicable to EPA mission goals. When a potential pollutant enters the environment, it will almost always become involved in chemical transformations that influence its movement through and its impact on the environment. Photochemical and biological factors may be active. The EPA should encourage systematic and fundamental research directed toward clarification of reaction pathways open to molecules, atoms, and ions of environmental interest, both in the gas phase and in aqueous solutions. More important, however, EPA should have as a conscious and publicized goal the detection of potentially undesirable environmental constituents at concentration levels far below known or expected toxicity limits. To reach this goal, EPA should stimulate the development of new analytical techniques of all kinds. Its program must, of course, include analytical detection of specific substances already known to present environmental issues. In addition, EPA should take a prominent role in the support of long-range research in analytical chemistry. Program emphases should include extension of sensitivity limits, increase in detection selectivity, and exploration of new concepts.

Next: Appendix A: Chemistry in Industry: Tables »
Opportunities in Chemistry Get This Book
×
Buy Hardback | $60.00
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

Opportunities in Chemistry is based on the contributions of hundreds of American chemists in academia and industry and should be taken as the best available consensus of the chemical community regarding its intellectual frontiers and the economic opportunities that lie beyond them," says Science. This volume addresses the direction in which today's chemical research is heading, including recent developments, technological applications, and the ways advances in chemistry can be used to improve the human condition. In addition, the book examines economic and political implications of chemical research and lists resources for basic research and education in the chemical sciences.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!