Cover Story  
  March 14,  2005
Volume 83, Number 11
pp. 41-45
 

PRIESTLEY MEDAL ADDRESS

  PROUD TO BE A CHEMIST  

   
 
 
GEORGE A. OLAH
PHOTO BY MITCH JACOBY
The title I chose emphasizes my lifetime love affair with chemistry and how proud I am to be a chemist.

I was born in Hungary in 1927 and entered the Technical University of my native Budapest in the fall of 1945 at the end of World War II. My father was a lawyer, and no one in my family had any interest in science. I was fortunate to have received a good general education with heavy emphasis on the humanities, history, languages, mathematics, and even some philosophy. While in school I never imagined becoming a scientist, least of all a chemist, but this changed when I took my first chemistry course at the university. I fell in love with chemistry, forming a lifelong attachment that still continues. I wrote more about my growing-up years, interests, etcetera, in my autobiographical book "A Life of Magic Chemistry."

I had some inspiring professors during my university years, including my future "doctor father" Geza Zemplen, a noted carbohydrate chemist of his time and a student of Emil Fischer's. (I am thus a scientific grandson of Fischer's, as noted on his "scientific family tree" published on the occasion of the centennial of his 1902 Nobel Prize.) Organic chemistry was taught at that time entirely in a descriptive manner, neglecting structural aspects and the emerging electronic theory. It was thus a challenge when, in 1953–54 as a young faculty member at the Technical University, I had the opportunity to teach a course titled "Introduction to Theoretical Organic Chemistry."

Some of my work with Zemplen on digitalis heart glycosides resulted in a practical process that was used for many years in Hungary to produce the heart medication digoxin. The starting point was the colorful foxglove plant, which thrived around Lake Balaton. Soon, however, my interests carried me in other directions. Leaving the shadow of Emil Fischer, I became interested in the developing area of organofluorine chemistry. Despite the difficult conditions and our limited research facilities in the post-WWII era, I developed new synthetic methods for the preparation of varied organofluorine compounds. By necessity, we needed to make our own HF from fluorspar and to make fluorosulfuric acid, boron trifluoride, and others as well. With no plastic equipment available, we distilled HF, for example, in a silver apparatus. Hans Meerwein, who read some of my early papers and sympathized with our difficulties, arranged to have a BF3 gas cylinder sent to me. What a magnificent gift it was! Incidentally, Meerwein always was my idol in chemistry. I was privileged to later meet him in person, and we maintained contact until his death in 1965. 

PRIESTLEY MEDAL ADDRESS

George A. Olah is scheduled to present the Priestley Medal Address on March 15 during the award ceremony at the American Chemical Society's 229th national meeting in San Diego. Olah, 77, is the 1994 Chemistry Nobel Laureate and director of the Loker Hydrocarbon Research Institute at the University of Southern California. He is known throughout the chemistry community for his groundbreaking work with hydrocarbons. Olah is being honored with the Priestley Medal for "his many extraordinarily creative and pioneering contributions to chemistry over a 50-year career and especially his recent insightful work on the methanol economy." Published here is the text of his address.

8311add_couple.tifcxd
IN PARTNERS The Olahs, shown here in 1996, have collaborated for some 55 years.

COURTESY OF GEORGE OLAH

IN BUDAPEST, I also cooperated with researchers at the Medical University to study the pharmaceutical--and particularly the anticancer--activity of some of our organofluorine compounds. These studies in the early 1950s were among the earliest in the field, which since has gained great significance. Having made a variety of organofluorine reagents, I also became interested in their reactivity in Friedel-Crafts and related reactions such as alkylation, acylation, and nitration. Subsequently, I also began studying their reaction mechanisms, including the intermediates formed with acid catalysts such as boron trifluoride. Thus started my fascination with electrophilic reactions and their intermediates.

Of all the difficult aspects of life in Communist-dominated Hungary, isolation from the world scientific community was the most depressing for researchers and teachers. Regardless, we tried to make the best of the existing limited possibilities. In 1954, I joined the new Chemical Research Institute of the Hungarian Academy of Sciences and had a small research group in organic chemistry.

In October 1956, Hungary revolted against Soviet rule, but the uprising was soon put down by drastic measures and with much loss of life. Budapest was devastated, and the future looked bleak. There was nearly unanimous support for the spontaneous revolt of 1956 in Budapest, but that support could not on its own prevent the oppression and terror that returned with the Soviet forces. With the Iron Curtain temporarily lowered, in November and early December 1956 some 200,000 Hungarians, mostly of the younger generation, fled their homeland.

My family (our older son, George, was born in 1954) and most of my research group, who also decided that this was the only path to follow, joined the torrent of refugees seeking a new life in the West.

Eventually, we got to London, where an aunt of my wife lived and where we were warmly welcomed. During our stay in London, I was able to meet for the first time some of the chemists whose work I knew from the literature and whom I admired. I found them most gracious and helpful. In particular, Christopher Ingold and Alexander Todd extended their efforts on behalf of a young, little-known Hungarian refugee chemist in a way that I will never forget and for which I will always be grateful. I also remember being invited in early January 1957 to give a seminar at Cambridge. This was the first time I ever lectured in English. My talk raised some comments such as, "How interesting it is that this strange Hungarian language has some words which resemble English." Anyhow, I survived my "baptism," but, like many Hungarian-born scientists, I retain to this day an unmistakable accent.

My family and I did not intend to settle in England and in March 1957 moved on to Montreal, where my mother-in-law lived, having remarried there after the war. I initially looked for an academic position, but none came along. A few industrial research possibilities opened up; of these, I accepted one with Dow Chemical. Dow, with its home base in Midland, Mich., was establishing a small exploratory research laboratory 100 miles away across the Canadian border in Sarnia, Ontario, where its Canadian subsidiaries' major operations were located. Two of my former Hungarian associates, including Steve Kuhn, joined me at Dow. I had a productive and rewarding time with Dow, first in Canada and subsequently at Dow's Eastern Research Laboratory in the Boston area.

It was during my "industrial period" in 1959–60 that my breakthrough work on persistent, long-lived carbocations was started. We found that in extremely strong acid systems (so-called superacids such as SbF5 or HSO3F-SbF5), alkyl halides, alcohols, and so on, readily ionize and form remarkably stable positive ions (subsequently called carbocations). This was the beginning of the study of stable (persistent) carbocations, which subsequently was greatly extended and involved many laboratories around the world.

The name "superacid," incidentally, was suggested by no lesser a chemist than J. B. Conant of Harvard. In a paper in 1927, he denoted as such acids (notably perchloric acid) that he found stronger than conventional mineral acids and that were able to protonate weak bases, including carbonyl compounds.

Until the late 1950s, however, chemists still considered mineral acids, sulfuric acid first of all, to be the strongest acid systems in existence. This idea started to change with the above-mentioned studies using extremely strong acid systems. They eventually turned out to be many billions or even trillions of times stronger than sulfuric acid. R. J. Gillespie--who also should be credited with the development of much of the fundamental inorganic chemistry of superacids--suggested in the early 1960s that protic acids stronger than 100% sulfuric acid should be classified as superacids. This arbitrary but most useful definition was generally accepted.

In our studies of long-lived carbocations, we also used strong Lewis acids--such as antimony pentafluoride, arsenic pentafluoride, and tantalum pentafluorides--so we also arbitrarily named Lewis acids stronger than anhydrous aluminum trichloride "Lewis superacids." The superacids most used in our studies, however, were the conjugate acids of fluorosulfuric acid, hydrogen fluoride, trifluoromethanesulfuric acid, and others with the strong Lewis acid fluorides. The widely used FSO3H-SbH5 system was nicknamed in our laboratory "magic acid," which later became its trade name.

During my years with Dow, my interest also extended to the broader study and use of superacids, particularly their applications in catalytic chemistry. I also extensively studied electrophilic aromatic substitutions and their intermediates, which are of practical interest, for example, in alkylations such as the ethylation of benzene for styrene production. In mechanistic studies of electrophilic nitration with stable nitronium salts, we were able to show that substrate and positional selectivities of reactive aromatic hydrocarbons are determined in separate steps. Other studies included those of new synthetic reagents and various mechanistic problems. I am grateful to Dow for having allowed me to pursue these studies and even to publish many results.

In 1965, at the recommendation of Paul D. Bartlett from Harvard, one of the leading academic organic chemists, I was able to return to university life as professor and chairman of the department of chemistry at Western Reserve University in Cleveland. I assume not many of Bartlett's former students or other acquaintances were much interested in this position at a small midwestern school. At different times, I myself had some contacts with the University of Toronto, Ohio State University, and Princeton University, but nothing came of them. Reluctance to take a chance on a young immigrant chemist from faraway Hungary who had been working in industry and without proper North American academic "pedigree" must have been a factor.

PRIDE & JOY In an Olah family portrait from 1962, Ron stands next to his mother, Judy, and George, next to his father, George. The next generation of Olahs poses more informally; (from left) Kaitlyn, Justin, and Peter.
PHOTOS COURTESY OF GEORGE OLAH

8311odd_boy1.tifcxd
COURTESY OF GEORGE OLAH
8311cov1_olah1.tifcxd
BEFORE & AFTER As a youngster, Olah was an avid reader--but not usually about science; now he's in his sixth decade as a scientist.

PHOTO BY MITCH JACOBY

NEVERTHELESS, my years in Cleveland were very productive and rewarding. In three years, we merged the neighboring chemistry departments of Case Institute of Technology and Western Reserve University, under my chairmanship, into a single, stronger department, realizing that such a unification would be to everyone's advantage. Our effort was so smooth and successful that it soon led to the complete merger of the two universities into Case Western Reserve University. This outcome was most rewarding, and feeling that I had made my contribution, I soon was able to give up my administrative duties. I always had wanted only to teach and do research and had "survived" my administrative role without changing my goals in any way.

The Olah group rapidly expanded as dedicated graduate students and postdoctoral fellows joined. Our research involved a number of areas including the continued exploration of the wide scope of superacid and carbocation chemistry. Stable (persistent) carbocation studies involved not only the usual trivalent but also five- and higher-coordination ions. The much-publicized, so-called nonclassical ion controversy centered mainly on the case of the norbornyl cation. The 1962 Brookhaven Mechanism Conference, where I first reported on our work on long-lived carbocations in public, is still clear in my mind. The scheduled "main event" of the meeting was the continuing debate between Saul Winstein and Herbert Brown on the classical or nonclassical nature of the norbornyl cation. It must have come as a surprise to them and to the audience that a young chemist from an industrial laboratory had been invited to lecture on having obtained and studied stable, long-lived carbonium ions (as they were still called at the time) by the new method of using highly acidic (superacidic) systems.

I remember being called aside separately during the conference by Winstein and by Brown, both towering and dominating personalities of the time, who cautioned me that a young man should be exceedingly careful in making such claims. Each pointed out that most probably I was wrong and could not have obtained long-lived carbonium ions. Just in case my method turned out to be real, however, I was advised to obtain evidence for the "nonclassical" or "classical" nature (depending on who was giving the advice) of the much-disputed 2-norbornyl cation. Eventually, the norbornyl ion turned out to be the -delocalized higher coordinate nonclassical ion and not a pair of equilibrating trivalent classical ions. This ended the dispute and a much-discussed chapter of physical-organic chemistry.

Although many people say too much effort was expended on the "futile" nonclassical ion controversy, I believe that it eventually resulted in significant insights and much new chemistry. Based on the extensive studies of carbocations, it was possible for me to generalize the concept. I also suggested naming all cations of carbon compounds as "carbocations" (because the corresponding anions were named "carbanions"). Carbocations could be differentiated into trivalent (classical) carbenium ions, in contrast to five or higher coordinate (nonclassical) carbonium ions.

The study of carbocations also affected in a fundamental way our understanding of the chemical bonding of carbon compounds. It allowed us to extend Kekule's concept of the limiting ability of carbon to bind with no more than four other atoms or groups to five (or even six or seven) atoms or groups. It also led to the realization of the general ability of saturated C–H and C–C 2e-2c Lewis-type single bonds to act via 2e-3c bonding as -donors toward varied strong electrophiles or highly reactive reagents in superacidic systems. The electrophilic chemistry of saturated hydrocarbons (including that of the parent methane) started to evolve based on the recognition of the concept and significance of hypercoordinate carbon (hypercarbon chemistry).

During my Cleveland years, we also investigated varied onium ion systems, developed new synthetic reagents and methods, continued investigations in organofluorine and organometallic chemistry, carried out systematic studies in Friedel-Crafts and related chemistry, explored new aspects of electrophilic reactions of aliphatic and aromatic hydrocarbons, and extended work on nitration chemistry, to mention some of the main areas.

Having achieved most of the goals I set out for myself in Cleveland, in 1977 I felt it was time for change, including that of my major research direction. Our younger son, Ron, was entering college at the time, and his heart was set on Stanford. He felt it would be nice if the whole family moved to California. Coincidentally, at the same time I had an offer from the University of Southern California. It was thus easy to agree, and so we moved to Los Angeles. With my friend Sid Benson, we started the Hydrocarbon Research Institute, which was subsequently named the Loker Hydrocarbon Research Institute in honor of Don and Katherine Loker, great friends and benefactors of the university and of our institute. I found a challenging new home at USC, which I still much enjoy.

I have been interested during my entire career in the chemistry of hydrocarbons. The oil crises of the 1970s put this field into a new perspective. Clearly, there is a need to develop not only new and more efficient ways to use still-existing natural hydrocarbon sources (oil and gas, as well as coal), but also new approaches to synthetic hydrocarbons other than by Fischer-Tropsch chemistry. Our institute is dedicated to these goals. At the same time, I continued some of my previous research interests, including studying carbocations and their chemistry, for which I was awarded the 1994 Nobel Prize in Chemistry. We extended the scope of new hypercarbon chemistry and started to develop the fascinating new area of superelectrophiles and superelectrophilic activation.

My work in Los Angeles, however, continues to center on the study of new hydrocarbon chemistry. We search for new approaches to produce and utilize hydrocarbons beyond the availability of nonrenewable and relatively cheap oil and gas. The major effort of our present research is directed toward developing an approach that I call the "methanol economy." The approach involves producing methanol directly from still-abundant vast natural gas resources (that is, methane) without going through syngas. More important, we also pursue new chemistry to reductively convert carbon dioxide to methanol.

8311odd_wedding.tifcxd
WEDDING DAY In 1949, newlyweds George and Judy Olah walk under oars held by members of Olah's rowing club.

COURTESY OF GEORGE OLAH

CARBON DIOXIDE can be readily separated from flue gases of coal-burning power plants or industrial plants. Rather than just sequestering CO2, the gas can be used for producing methanol-based fuels and raw materials for hydrocarbons. I believe that, using selective absorption methods and membrane technology, it will be eventually even feasible to separate atmospheric CO2 itself (representing only 0.036% of air) and convert it into methanol, thus freeing humankind of its reliance on diminishing fossil fuels. Of course, to produce the needed hydrogen by electrolysis of water, much energy is needed, which will be provided by atomic energy and the use of all alternative energy sources. Other approaches (photocatalytic, enzymatic, etcetera) also offer possibilities.

Methanol, a convenient liquid, is a way to store and transport energy. Methanol is also an excellent fuel in its own right, including its use in the direct methanol fuel cell (DMFC) that we developed jointly with the Jet Propulsion Laboratory of California Institute of Technology. Methanol, significantly, can also be directly converted catalytically into ethylene or propylene and subsequently to varied hydrocarbons and their products presently obtained from oil and natural gas. Once it becomes economically feasible to chemically recycle atmospheric CO2, the process will also supplement nature's photosynthesis to mitigate the effect of this major greenhouse gas on global warming. We are involved in extensive research on all these issues. I believe that this is probably the most interesting and significant work I have ever been involved in and that it has real viability. In contrast to the ease of handling liquid methanol, other approaches--such as the much-discussed hydrogen economy--must handle and transport an extremely volatile, potentially explosive gas under pressure, necessitating a new and very expensive infrastructure. I am not suggesting that we don't pursue all possible avenues to solve our dependence on diminishing fossil fuels. The methanol economy approach, however, certainly deserves serious consideration as well as further research and evaluation.

It was always my practice during my career that whenever I felt that I had substantially achieved my goals and fulfilled my interest in a specific area of research, I wrote (or edited) a comprehensive monograph on the field. Those interested can consult these books for details.

I am most grateful for the contribution of all my former students, postdoctoral researchers, and other associates--by now they exceed 200--who made our joint work possible and whose names are to be found on our publications. I always had an open-door policy and was available to my students and associates, whom I consider members of my broader scientific family. With my wife, Judy, we keep in contact with them as much as possible, not only scientifically but also at the personal level. Special recognition and thanks are due to my friend, former student, and longtime colleague professor Surya Prakash, who in recent years, besides his own independent work, increasingly plays a major role in our joint research programs.

I have spent a lifetime in research and teaching of chemistry, which I always considered a bridge between the different branches of science (such as physics and biology). I discussed in my autobiography my views on chemistry as the multifaceted central science in more detail, including the historical aspects of how chemistry developed and where it is heading. Inevitably, science has become more specialized and subdivided, but I strongly believe in the universality of science. Whereas interdisciplinary study and research are essential and provide much of the most significant and intriguing advances, they cannot and should not be pursued without building on solid fundamentals. There can be no substitute to first acquiring solid foundations in the basic underlying disciplines of mathematics, physics, and chemistry. Would a high-rise building be stable if construction started at its higher levels, lacking solid foundations? 

THE WRITTEN WORD
George Olah Is A Prolific Author, Editor

AUTOBIOGRAPHICAL

"A Life of Magic Chemistry," Wiley, 2001.

"Crossing conventional lines in 50 years of my research," Journal of Organic Chemistry, upcoming 2005.

SCIENTIFIC

(John Wiley & Sons is publisher unless otherwise indicated.)

"Introduction to Theoretical Organic Chemistry" (in German), Akad. Verlag, 1960.

"Friedel-Crafts and Related Reactions," Vols. I–IV (ed.), 1963–64.

"Carbonium Ions" (ed. with Paul Schleyer), Vols. I–V, 1968–1972.

"Friedel-Crafts Chemistry," 1973.

"Carbocations and Electrophilic Reactions," 1973.

"Halonium Ions," 1975.

"Superacids" (with Surya Prakash and Jean Sommer), 1985.

"Hypercarbon Chemistry" (with Surya Prakash, Robert Williams, Leslie Field, and Kenneth Wade), 1987.

"Nitration" (with Ripumadan Malhotra and Subhash Narang), VCH, 1989.

"Cage Hydrocarbons" (ed.), 1990.

"Chemistry of Energetic Materials" (ed. with David Squire), Academic Press, 1991.

"Electron-Deficient Boron and Carbon Clusters" (ed. with Kenneth Wade and Robert Williams), 1991.

"Synthetic Fluorine Chemistry" (ed. with Richard Chambers and Surya Prakash), 1992.

"Hydrocarbon Chemistry" (with Arpad Molnar), 1995.

"Onium Ions" (with Kenneth Laali, Qi Wang, and Surya Prakash), 1998.

"Across Conventional Lines, selected papers of George Olah" (ed. with Surya Prakash), World Scientific Publishing, Singapore, 2003.

"Carbocation Chemistry" (ed. with Surya Prakash), Wiley-Interscience, 2004.

"Beyond Oil and Gas: The Methanol Economy" (with Alain Goeppert and Surya Prakash), Wiley-VCH, 2005 (in preparation).

SOME OF today's often-discussed trends are disturbing. The question of the survivability of chemistry as a field in its own right has even been raised. Many people see chemists only as general molecular scientists or engineers. Of course, chemistry is essentially a broad-scope molecular science--we are dealing with molecules and their assemblies--but it is built on its own common foundations. I fully appreciate the great significance of biological chemistry, the chemistry of materials science, and other interdisciplinary fields in which chemistry plays a fundamental role. However, we as chemists are contributing in a major way to these fields.

I am passionate about science and proud to be a chemist. During my career I always crossed conventional narrow boundaries, which came naturally to me. I started as a natural product carbohydrate chemist. My subsequent contributions, mainly in synthetic and mechanistic organic chemistry, were frequently based on the use and development of new inorganic reagents and systems (such as superacids). I also always tried to take full advantage of emerging new physical spectroscopic methods. After some reluctance (although during my Cleveland years we had regular joint research group meetings with John Pople and Ned Arnett from Pittsburgh), I also adopted with my associates the powerful tools of computational methods, mostly at the urging of my friend Paul Schleyer. I never, however, considered myself a specialized physical-organic or any other "hyphenated chemist," just a chemist. In my research and teaching I always tried to emphasize the general common principles and universality of chemistry. I hope I have also succeeded in inspiring my students and younger colleagues along these lines.

Reflecting on my life, a few years ago I wrote the following in my autobiography: "My long journey, which started in my native Budapest on the banks of the Danube and took me to the shores of the Pacific Ocean, was not always an easy one. Human nature, however, helps to block out memories of hardship and difficulties. They fade away and you look back remembering mostly the positive aspects of your life. I followed my own principles and went my own way. It helped that I inherited a strong, perhaps on occasion stubborn, nature with a determination to follow the pathway to my goals and that I worked hard to achieve them. It was and still is a rewarding life, which I share with my wife, Judy, including our common profession." Judy's love, understanding for my many shortcomings, and her strength to keep our family on course cannot be thanked in simple words. Our son George is an M.B.A. and the chief financial officer of a financial company in Los Angeles; Ron, a physician, is practicing internal medicine in Pasadena. We have two wonderful daughters-in-law, Sally and Cindy, and three adorable grandchildren, Peter, Kaitlyn, and Justin.?

I would like to end by thanking my adopted country, which welcomed us as immigrants more than four decades ago. I firmly believe in America; it still offers a new home and practically unlimited possibilities to the newcomers who are willing to work hard for them. It is also where much of the main action in science and technology remains. I hope that the "American dream," which we were so fortunate to attain, will remain a reality for future generations.

MORE ON THIS COVER STORY

PRIESTLEY MEDALIST
A Love Affair With Chemistry
Nobel Laureate George A. Olah has spent nearly six decades in the service of chemistry
 

 
     
  Chemical & Engineering News
ISSN 0009-2347
Copyright © 2005