1976 & 1992 CHEMISTRY NOBEL PRIZE AUTOGRAPH MARCUS LIPSCOMB BORANES ELECTRONS For Sale

1976 & 1992 CHEMISTRY NOBEL PRIZE AUTOGRAPH MARCUS LIPSCOMB BORANES ELECTRONS
When you click on links to various merchants on this site and make a purchase, this can result in this site earning a commission. Affiliate programs and affiliations include, but are not limited to, the eBay Partner Network.


Buy Now

1976 & 1992 CHEMISTRY NOBEL PRIZE AUTOGRAPH MARCUS LIPSCOMB BORANES ELECTRONS:
$246.45

TWO 4X6 SIGNED TYPED QUOTES SIGNED INDEX CARDS WITH WRITING ON WHY THEY WON THE NOBEL PRIZED IN CHEMISTRY... SIGNED BY RUDOLPH A. MARCUS AND WILLIAM N. LIPSCOMB JR.Rudolf Marcus is a theoretical chemist who studied electron transfer in redox (reduction and oxidation) reactions. Rudolf established what is now known as the Marcus theory to explain electron transfer reactions. Rudolf was awarded the 1992 Nobel Prize in Chemistry for this important research.
Electron transfer reactions describe the rate at which an electron can jump from one chemical species to another. Rudolf’s work showed how molecular changes in reactants and solvent molecules influence this intermolecular movement of electrons. Marcus theory is able to describe important chemical and biological processes, including photosynthesis, corrosion, types of chemiluminescence — the emission of light resulting from a chemical reaction — and charge separation in some types of solar cell.
The interaction between experiment and theory, each stimulating the other, remains one of the joys of Rudolf\'s scientific experience. He draws upon the experimental findings or puzzles of others to uncover theoretical problems to study whilst his own practical research has flavoured his attitude towards and interest in theoretical research.
y first encounters with McGill University came when I was still in a baby carriage. My mother used to wheel me about the campus when we lived in that neighborhood and, as she recounted years later, she would tell me that I would go to McGill. There was some precedent for my going there, since two of my father’s brothers received their M.D.’s at McGill.
I have always loved going to school. Since neither of my parents had a higher education, my academic “idols” were these two paternal uncles and one of their uncles, my great-uncle, Henrik Steen (né Markus). My admiration for him, living in faraway Sweden, was not because of a teol.dr. (which he received from the University of Uppsala in 1915) nor because of the many books he wrote – I knew nothing of that – but rather because he was reputed to speak 13 languages. I learned decades later that the number was only 9! Growing up, mostly in Montreal, I was an only child of loving parents. I admired my father’s athletic prowess – he excelled in several sports – and my mother’s expressive singing and piano playing.
My interest in the sciences started with mathematics in the very beginning, and later with chemistry in early high school and the proverbial home chemistry set. My education at Baron Byng High School was excellent, with dedicated masters (boys and girls were separate). I spent the next years at McGill University, for both undergraduate and, as was the custom of the time, graduate study. Our graduate supervisor, Carl A. Winkler, specialized in rates of chemical reactions. He himself had received his Ph.D. as a student of Cyril Hinshelwood at Oxford. Hinshelwood was later the recipient of the Nobel Prize for his work on chemical kinetics. Winkler brought to his laboratory an enthusiastic joyousness in research and was much loved by his students.
During my McGill years, I took a number of math courses, more than other students in chemistry. Upon receiving a Ph.D. from McGill University in 1946, I joined the new post-doctoral program at the National Research Council of Canada in Ottawa. This program at NRC later became famous, but at the time it was still in its infancy and our titles were Junior Research Officers. The photochemistry group was headed by E.W.R. Steacie, an international figure in the study of free-radical reactions and a major force in the development of the basic research program at NRC. I benefitted from the quality of his research on gas phase reaction rates. Like my research on chemical reaction rates in solution at McGill (kinetics of nitration), it was experimental in nature. There were no theoretical chemists in Canada at the time, and as students I don’t think we ever considered how or where theories were conceived.
About 1948 a fellow post-doctoral at NRC, Walter Trost, and I formed a two-man seminar to study theoretical papers related to our experimental work. This adventure led me to explore the possibility of going on a second post-doctoral, but in theoretical work, which seemed like a radical step at the time. I had a tendency to break the glass vacuum apparatus, due to a still present impetuous haste, with time-consuming consequences. Nevertheless, the realization that breaking a pencil point would have far less disastrous consequences played little or no role, I believe, in this decision to explore theory!
I applied in 1948 to six well-known theoreticians in the U.S. for a postdoctoral research fellowship. The possibility that one of them might take on an untested applicant, an applicant hardly qualified for theoretical research, was probably too much to hope for. Oscar K. Rice at the University of North Carolina alone responded favorably, subject to the success of an application he would make to the Office of Naval Research for this purpose. It was, and in February 1949 I took the train south, heading for the University of North Carolina in Chapel Hill. I was impressed on arrival there by the red clay, the sandy walks, and the graciousness of the people.
After that, I never looked back. Being exposed to theory, stimulated by a basic love of concepts and mathematics, was a marvelous experience. During the first three months I read everything I could lay my hands on regarding reaction rate theory, including Marcelin’s classic 1915 theory which came within one small step of the Transition State Theory of 1935. I read numerous theoretical papers in German, a primary language for the “chemical dynamics” field in the 1920s and 1930s, attended my first formal course in quantum mechanics, given by Nathan Rosen in the Physics Department, and was guided by Oscar in a two-man weekly seminar in which I described a paper I had read and he pointed out assumptions in it that I had overlooked. My life as a working theorist began three months after this preliminary study and background reading, when Oscar gently nudged me toward working on a particular problem.
Fortunately for me, Oscar’s gamble paid off. Some three months later, I had formulated a particular case of what was later entitled by B. Seymour Rabinovitch, RRKM theory (“Rice-Ramsperger-Kassel-Marcus”). In it, I blended statistical ideas from the RRK theory of the 1920s with those of the transition state theory of the mid-1930s. The work was published in 1951. In 1952 I wrote the generalization of it for other reactions. In addition, six months after arrival in Chapel Hill, I was also blessed by marriage to Laura Hearne, an attractive graduate student in sociology at UNC. She is here with me at this ceremony. Our three sons, Alan, Kenneth and Raymond, and two daughters-in-law are also present today.
In 1951, I attempted to secure a faculty position. This effort met with little success (35 letters did not yield 35 no’s, since not everyone replied!). Very fortunately, that spring I met Dean Raymond Kirk of the Polytechnic Institute of Brooklyn at an American Chemical Society meeting in Cleveland, which I was attending primarily to seek a faculty position. This meeting with Dean Kirk, so vital for my subsequent career, was arranged by Seymour Yolles, a graduate student at UNC in a course I taught during Rice’s illness. Seymour had been a student at Brooklyn Poly and learned, upon accidentally encountering Dr. Kirk, that Kirk was seeking new faculty. After a subsequent interview at Brooklyn Poly, I was hired, and life as a fully independent researcher began.
I undertook an experimental research program on both gas phase and solution reaction rates, wrote the 1952 RRKM papers, and wondered what to do next in theoretical research. I felt at the time that it was pointless to continue with RRKM since few experimental data were available. Some of our experiments were intended to produce more.
After some minor pieces of theoretical study that I worked on, a student in my statistical mechanics class brought to my attention a problem in polyelectrolytes. Reading everything I could about electrostatics, I wrote two papers on that topic in 1954/55. This electrostatics background made me fully ready in 1955 to treat a problem I had just read about on electron transfers. I comment on this next period on electron transfer research in my Nobel Lecture. About 1960, it became clear that it was best for me to bring the experimental part of my research program to a close – there was too much to do on the theoretical aspects – and I began the process of winding down the experiments. I spent a year and a half during 1960-61 at the Courant Mathematical Institute at New York University, auditing many courses which were, in part, beyond me, but which were, nevertheless, highly instructive.
In 1964, I joined the faculty of the University of Illinois in Urbana-Champaign and I never undertook any further experiments there. At Illinois, my interests in electron transfer continued, together with interests in other aspects of reaction dynamics, including designing “natural collision coordinates”, learning about action-angle variables, introducing the latter into molecular collisions, reaction dynamics, and later into semiclassical theories of collisions and of bound states, and spending much of my free time in the astronomy library learning more about classical mechanics, celestial mechanics, quasiperiodic motion, and chaos. I spent the academic year of 1975-76 in Europe, first as Visiting Professor at the University of Oxford and later as a Humboldt Awardee at the Technical University of Munich, where I was first exposed to the problem of electron transfer in photosynthesis.
In 1978, I accepted an offer from the California Institute of Technology to come there as the Arthur Amos Noyes Professor of Chemistry. My semiclassical interlude of 1970-80 was intellectually a very stimulating one, but it involved for me less interaction with experiments than had my earlier work on unimolecular reaction rates or on electron transfers. Accordingly, prompted by the extensive experimental work of my colleagues at Caltech in these fields of unimolecular reactions, intramolecular dynamics and of electron transfer processes, as well as by the rapidly growing experimental work in both broad areas world-wide, I turned once again to those particular topics and to the many new types of studies that were being made. Their scope and challenge continues to grow to this day in both fields. Life would be indeed easier if the experimentalists would only pause for a little while!
There was a time when I had wondered about how much time and energy had been lost doing experiments during most of my stay at Brooklyn Poly- experiments on gas phase reactions, flash photolysis, isotopic exchange electron transfer, bipolar electrolytes, nitration, and photoelectrochemistry, among others-and during all of my stay at NRC and at McGill. In retrospect, I realized that this experimental background heavily flavored my attitude and interests in theoretical research. In the latter I drew, in most but not all cases, upon experimental findings or puzzles for theoretical problems to study. The growth of experiments in these fields has served as a continually rejuvenating influence. This interaction of experiment and theory, each stimulating the other, has been and continues to be one of the joys of my experience.
Honors received for the theoretical work include the Irving Langmair and the Peter Debye Awards of the American Chemical Society (1978, 1988), the Willard Gibbs, Theodore William Richards, and Pauling Medals, and the Remsen and Edgar Fahs Smith Awards, from various sections of the ACS, (1988, 1990, 1991, 1991, 1991), the Robinson and the Centenary Medals of the Faraday Division of the Royal Society of Chemistry (1982, 1988), Columbia University’s Chandler Medal (1983) and Ohio State’s William Lloyd Evans Award (1990), a Professorial Fellowship at University College, Oxford (1975 to 1976) and a Visiting Professorship in Theoretical Chemistry at Oxford during that period, the Wolf Prize in Chemistry (1985), the National Medal of Science (1989), the Hirschfelder Prize in Chemistry (1993), election to the National Academy of Sciences (1970), the American Academy of Arts and Sciences (1973), the American Philosophical Society (1990), honorary membership in the Royal Society of Chemistry (1991), and foreign membership in the Royal Society (London) (1987) and in the Royal Society of Canada (1993). Honorary degrees were conferred by the University of Chicago and by Goteborg, Polytechnic, McGill, and Queen’s Universities and by the University of New Brunswick (1983, 1986, 1987, 1988, 1993, 1993). A commemorative issue of the Journal of Physical Chemistry was published in 1986.
William N. LipscombThe Nobel Prize in Chemistry 1976
Born: 9 December 1919, Cleveland, OH, USA
Died: 14 April 2011, Cambridge, MA, USA
Affiliation at the time of the award: Harvard University, Cambridge, MA, USA
Prize motivation: \"for his studies on the structure of boranes illuminating problems of chemical bonding.\"
William Nunn Lipscomb Jr. (December 9, 1919 – April 14, 2011)[2] was a Nobel Prize-winning American inorganic and organic chemist working in nuclear magnetic resonance, theoretical chemistry, boron chemistry, and biochemistry.Contents1 Biography1.1 Overview1.2 Early years1.3 Education1.4 Later years2 Scientific studies2.1 Nuclear magnetic resonance and the chemical shift2.2 Boron chemistry and the nature of the chemical bond2.3 Large biological molecule structure and function2.4 Other results3 Positions, awards and honors4 References5 External linksBiographyOverviewLipscomb was born in Cleveland, Ohio. His family moved to Lexington, Kentucky in 1920,[1] and he lived there until he received his Bachelor of Science degree in Chemistry at the University of Kentucky in 1941. He went on to earn his Doctor of Philosophy degree in Chemistry from the California Institute of Technology (Caltech) in 1946.
From 1946 to 1959 he taught at the University of Minnesota. From 1959 to 1990 he was a professor of chemistry at Harvard University, where he was a professor emeritus since 1990.
Lipscomb was married to the former Mary Adele Sargent from 1944 to 1983.[3] They had three children, one of whom lived only a few hours. He married Jean Evans in 1983.[4] They had one adopted daughter.
Lipscomb resided in Cambridge, Massachusetts until his death in 2011 from pneumonia.[5]
Early years\"My early home environment ... stressed personal responsibility and self reliance. Independence was encouraged especially in the early years when my mother taught music and when my father\'s medical practice occupied most of his time.\"
In grade school Lipscomb collected animals, insects, pets, rocks, and minerals.
Interest in astronomy led him to visitor nights at the Observatory of the University of Kentucky, where Prof. H. H. Downing gave him a copy of Baker\'s Astronomy. Lipscomb credits gaining many intuitive physics concepts from this book and from his conversations with Downing, who became Lipscomb\'s lifelong friend.
The young Lipscomb participated in other projects, such as Morse-coded messages over wires and crystal radio sets, with five nearby friends who became physicists, physicians, and an engineer.
At age of 12, Lipscomb was given a small Gilbert chemistry set. He expanded it by ordering apparatus and chemicals from suppliers and by using his father\'s privilege as a physician to purchase chemicals at the local drugstore at a discount. Lipscomb made his own fireworks and entertained visitors with color changes, odors, and explosions. His mother questioned his home chemistry hobby only once, when he attempted to isolate a large amount of urea from urine.
Lipscomb credits perusing the large medical texts in his physician father\'s library and the influence of Linus Pauling years later to his undertaking biochemical studies in his later years. Had Lipscomb become a physician like his father, he would have been the fourth physician in a row along the Lipscomb male line.
The source for this subsection, except as noted, is Lipscomb\'s autobiographical sketch.[6]
EducationLipscomb\'s high-school chemistry teacher, Frederick Jones, gave Lipscomb his college books on organic, analytical, and general chemistry, and asked only that Lipscomb take the examinations. During the class lectures, Lipscomb in the back of the classroom did research that he thought was original (but he later found was not): the preparation of hydrogen from sodium formate (or sodium oxalate) and sodium hydroxide.[7] He took care to include gas analyses and to search for probable side reactions.
Lipscomb later had a high-school physics course and took first prize in the state contest on that subject. He also became very interested in special relativity.
In college at the University of Kentucky Lipscomb had a music scholarship. He pursued independent study there, reading Dushman\' s Elements of Quantum Mechanics, the University of Pittsburgh Physics Staff\'s An Outline of Atomic Physics, and Pauling\'s The Nature of the Chemical Bond and the Structure of Molecules and Crystals. Prof. Robert H. Baker suggested that Lipscomb research the direct preparation of derivatives of alcohols from dilute aqueous solution without first separating the alcohol and water, which led to Lipscomb\'s first publication.[8]
For graduate school Lipscomb chose Caltech, which offered him a teaching assistantship in Physics at $20/month. He turned down more money from Northwestern University, which offered a research assistantship at $150/month. Columbia University rejected Lipscomb\'s application in a letter written by Nobel prizewinner Prof. Harold Urey.
At Caltech Lipscomb intended to study theoretical quantum mechanics with Prof. W. V. Houston in the Physics Department, but after one semester switched to the Chemistry Department under the influence of Prof. Linus Pauling. World War II work divided Lipscomb\'s time in graduate school beyond his other thesis work, as he partly analyzed smoke particle size, but mostly worked with nitroglycerin–nitrocellulose propellants, which involved handling vials of pure nitroglycerin on many occasions. Brief audio clips by Lipscomb about his war work may be found from the External Links section at the bottom of this page, past the References.
The source for this subsection, except as noted, is Lipscomb\'s autobiographical sketch.[6]
Later yearsThe Colonel is how Lipscomb\'s students referred to him, directly addressing him as Colonel. \"His first doctoral student, Murray Vernon King, pinned the label on him, and it was quickly adopted by other students, who wanted to use an appellation that showed informal respect. ... Lipscomb\'s Kentucky origins as the rationale for the designation.\"[9] Some years later in 1973 Lipscomb was made a member of the Honorable Order of Kentucky Colonels.[10]
Lipscomb, along with several other Nobel laureates, was a regular presenter at the annual Ig Nobel Awards Ceremony, last doing so on September 30, 2010.[11][12]
Scientific studiesLipscomb has worked in three main areas, nuclear magnetic resonance and the chemical shift, boron chemistry and the nature of the chemical bond, and large biochemical molecules. These areas overlap in time and share some scientific techniques. In at least the first two of these areas Lipscomb gave himself a big challenge likely to fail, and then plotted a course of intermediate goals.
Nuclear magnetic resonance and the chemical shift
NMR spectrum of hexaborane B6H10 showing the interpretation of a spectrum to deduce the molecular structure. (click to read details)In this area Lipscomb proposed that: \"... progress in structure determination, for new polyborane species and for substituted boranes and carboranes, would be greatly accelerated if the [boron-11] nuclear magnetic resonance spectra, rather than X-ray diffraction, could be used.\"[13] This goal was partially achieved, although X-ray diffraction is still necessary to determine many such atomic structures. The diagram at right shows a typical nuclear magnetic resonance (NMR) spectrum of a borane molecule.
Lipscomb investigated, \"... the carboranes, C2B10H12, and the sites of electrophilic attack on these compounds[14] using nuclear magnetic resonance (NMR) spectroscopy. This work led to [Lipscomb\'s publication of a comprehensive] theory of chemical shifts.[15] The calculations provided the first accurate values for the constants that describe the behavior of several types of molecules in magnetic or electric fields.\"[16]
Much of this work is summarized in a book by Gareth Eaton and William Lipscomb, NMR Studies of Boron Hydrides and Related Compounds,[17] one of Lipscomb\'s two books.
Boron chemistry and the nature of the chemical bondIn this area Lipscomb originally intended a more ambitious project: \"My original intention in the late 1940s was to spend a few years understanding the boranes, and then to discover a systematic valence description of the vast numbers of electron deficient intermetallic compounds. I have made little progress toward this latter objective. Instead, the field of boron chemistry has grown enormously, and a systematic understanding of some of its complexities has now begun.\"[18] Examples of these intermetallic compounds are KHg13 and Cu5Zn7. Of perhaps 24,000 of such compounds the structures of only 4,000 are known (in 2005) and we cannot predict structures for the others, because we do not sufficiently understand the nature of the chemical bond. This study was not successful, in part because the calculation time required for intermetallic compounds was out of reach in the 1960s, but intermediate goals involving boron bonding were achieved, sufficient to be awarded a Nobel Prize.
Lipscomb confirmed the molecular structure of boranes (compounds of boron and hydrogen) using X-ray crystallography in the 1950s and developed theories to explain their bonds. Later he applied the same methods to related problems, including the structure of carboranes (compounds of carbon, boron, and hydrogen).Atomic diagram of diborane (B2H6).
Bonding diagram of diborane (B2H6) showing with curved lines a pair of three-center two-electron bonds, each of which consists of a pair of electrons bonding three atoms, two boron atoms and a hydrogen atom in the middle.The three-center two-electron bond is illustrated in diborane (diagrams at right). In an ordinary covalent bond a pair of electrons bonds two atoms together, one at either end of the bond, the diboare B-H bonds for example at the left and right in the illustrations. In three-center two-electron bond a pair of electrons bonds three atoms (a boron atom at either end and a hydrogen atom in the middle), the diborane B-H-B bonds for example at the top and bottom of the illustrations.
Lipscomb\'s group did not propose or discover the three-center two-electron bond, nor did they develop formulas that give the proposed mechanism. In 1943, Longuet-Higgins, while still an undergraduate at Oxford, was the first to explain the structure and bonding of the boron hydrides. The paper reporting the work, written with his tutor R. P. Bell, [19] also reviews the history of the subject beginning with the work of Dilthey. [20] Shortly after, experimental spectroscopic work was performed by Price[21][22] that confirmed Longuet-Higgins\' structure for diborane. Longuet-Higgins and Roberts[23][24] discussed the electronic structure of an icosahedron of boron atoms and of the borides MB6. The mechanism of the three-center two-electron bond was also discussed in a later paper by Longuet-Higgins,[25] and an essentially equivalent mechanism was proposed by Eberhardt, Crawford, and Lipscomb.[26] Lipscomb\'s group also achieved an understanding of it through electron orbital calculations using formulas by Edmiston and Ruedenberg and by Boys.[27]
The Eberhardt, Crawford, and Lipscomb paper[26] discussed above also devised the \"styx number\" method to catalog certain kinds of boron-hydride bonding configurations.Diamond-square-diamond (DSD) rearrangement. At each vertex is a boron atom and (not shown) a hydrogen atom. A bond joining two triangular faces breaks to form a square, and then a new bond forms across opposite vertices of the square.Wandering atoms was a puzzle solved by Lipscomb[28] in one of his few papers with no co-authors. Compounds of boron and hydrogen tend to form closed cage structures. Sometimes the atoms at the vertices of these cages move substantial distances with respect to each other. The diamond-square-diamond mechanism (diagram at left) was suggested by Lipscomb to explain this rearrangement of vertices. Following along in the diagram at left for example in the faces shaded in blue, a pair of triangular faces has a left-right diamond shape. First, the bond common to these adjacent triangles breaks, forming a square, and then the square collapses back to an up-down diamond shape by bonding the atoms that were not bonded before. Other researchers have discovered more about these rearrangements.[29] [30]B10H16 showing in the middle a bond directly between two boron atoms without terminal hydrogens, a feature not previously seen in other boron hydrides.The B10H16 structure (diagram at right) determined by Grimes, Wang, Lewin, and Lipscomb found a bond directly between two boron atoms without terminal hydrogens, a feature not previously seen in other boron hydrides.[31]
Lipscomb\'s group developed calculation methods, both empirical[17] and from quantum mechanical theory.[32][33] Calculations by these methods produced accurate Hartree–Fock self-consistent field (SCF) molecular orbitals and were used to study boranes and carboranes.Ethane barrier to rotation about the carbon-carbon bond, first accurately calculated by Pitzer and Lipscomb.The ethane barrier to rotation (diagram at left) was first calculated accurately by Pitzer and Lipscomb[34] using the Hartree–Fock (SCF) method.
Lipscomb\'s calculations continued to a detailed examination of partial bonding through \"... theoretical studies of multicentered chemical bonds including both delocalized and localized molecular orbitals.\"[13] This included \"... proposed molecular orbital descriptions in which the bonding electrons are delocalized over the whole molecule.\"[35]
\"Lipscomb and his coworkers developed the idea of transferability of atomic properties, by which approximate theories for complex molecules are developed from more exact calculations for simpler but chemically related molecules,...\"[35]
Subsequent Nobel Prize winner Roald Hoffmann was a doctoral student [36] [37] [38] [39] [40] in Lipscomb\'s laboratory. Under Lipscomb\'s direction the Extended Hückel method of molecular orbital calculation was developed by Lawrence Lohr[18] and by Roald Hoffmann.[37][41] This method was later extended by Hoffman.[42] In Lipscomb\'s laboratory this method was reconciled with self-consistent field (SCF) theory by Newton[43] and by Boer.[44]
Noted boron chemist M. Frederick Hawthorne conducted early[45][46] and continuing[47][48] research with Lipscomb.
Much of this work is summarized in a book by Lipscomb, Boron Hydrides,[41] one of Lipscomb\'s two books.
The 1976 Nobel Prize in Chemistry was awarded to Lipscomb \"for his studies on the structure of boranes illuminating problems of chemical bonding\".[49] In a way this continued work on the nature of the chemical bond by his Doctoral Advisor at the California Institute of Technology, Linus Pauling, who was awarded the 1954 Nobel Prize in Chemistry \"for his research into the nature of the chemical bond and its application to the elucidation of the structure of complex substances.\"[50]
The source for about half of this section is Lipscomb\'s Nobel Lecture.[13][18]
Large biological molecule structure and functionLipscomb\'s later research focused on the atomic structure of proteins, particularly how enzymes work. His group used x-ray diffraction to solve the three-dimensional structure of these proteins to atomic resolution, and then to analyze the atomic detail of how the molecules work.
The images below are of Lipscomb\'s structures from the Protein Data Bank[51] displayed in simplified form with atomic detail suppressed. Proteins are chains of amino acids, and the continuous ribbon shows the trace of the chain with, for example, several amino acids for each turn of a helix.
carboxypeptidase Acarboxypeptidase ACarboxypeptidase A[52] (left) was the first protein structure from Lipscomb\'s group. Carboxypeptidase A is a digestive enzyme, a protein that digests other proteins. It is made in the pancreas and transported in inactive form to the intestines where it is activated. Carboxypeptidase A digests by chopping off certain amino acids one-by-one from one end of a protein. The size of this structure was ambitious. Carboxypeptidase A was a much larger molecule than anything solved previously.
apartate carbamoyltransferaseaspartate carbamoyltransferaseAspartate carbamoyltransferase.[53] (right) was the second protein structure from Lipscomb\'s group. For a copy of DNA to be made, a duplicate set of its nucleotides is required. Aspartate carbamoyltransferase performs a step in building the pyrimidine nucleotides (cytosine and thymidine). Aspartate carbamoyltransferase also ensures that just the right amount of pyrimidine nucleotides is available, as activator and inhibitor molecules attach to aspartate carbamoyltransferase to speed it up and to slow it down. Aspartate carbamoyltransferase is a complex of twelve molecules. Six large catalytic molecules in the interior do the work, and six small regulatory molecules on the outside control how fast the catalytic units work. The size of this structure was ambitious. Aspartate carbamoyltransferase was a much larger molecule than anything solved previously.
leucine aminopeptidaseLeucine aminopeptidaseLeucine aminopeptidase,[54] (left) a little like carboxypeptidase A, chops off certain amino acids one-by-one from one end of a protein or peptide.
HaeIII methyltransferaseHaeIII methyltransferase convalently complexed to DNAHaeIII methyltransferase[55] (right) binds to DNA where it methylates (adds a methy group to) it.
human interferon betahuman interferon betaHuman interferon beta[56] (left) is released by lymphocytes in response to pathogens to trigger the immune system.
chorismate mutasechorismate mutaseChorismate mutase[57] (right) catalyzes (speeds up) the production of the amino acids phenylalanine and (left) and its inhibitor MB06322 (CS-917)[59] were studied by Lipscomb\'s group in a collaboration, which included Metabasis Therapeutics, Inc., acquired by Ligand Pharmaceuticals[60] in 2010, exploring the possibility of finding a treatment for type 2 diabetes, as the MB06322 inhibitor slows the production of sugar by group also contributed to an understanding of concanavalin A[61] (low resolution structure), glucagon,[62] and carbonic anhydrase[63] (theoretical studies).
Subsequent Nobel Prize winner Thomas A. Steitz was a doctoral student in Lipscomb\'s laboratory. Under Lipscomb\'s direction, after the training task of determining the structure of the small molecule methyl ethylene phosphate,[64] Steitz made contributions to determining the atomic structures of carboxypeptidase A [52] [65] [66] [67] [68] [69] [70] [71] and aspartate carbamoyltransferase. [72] Steitz was awarded the 2009 Nobel Prize in Chemistry for determining the even larger structure of the large 50S ribosomal subunit, leading to an understanding of possible medical treatments.
Subsequent Nobel Prize winner Ada Yonath, who shared the 2009 Nobel Prize in Chemistry with Thomas A. Steitz and Venkatraman Ramakrishnan, spent some time in Lipscomb\'s lab where both she and Steitz were inspired to pursue later their own very large structures.[73] This was while she was a postdoctoral student at MIT in 1970.
Other results
Lipscombite: Mineral, small green crystals on quartz, Harvard Museum of Natural History, gift of W. N. Lipscomb Jr., 1996The mineral lipscombite (picture at right) was named after Professor Lipscomb by the mineralogist John Gruner who first made it artificially.
Low-temperature x-ray diffraction was pioneered in Lipscomb\'s laboratory[74][75][76] at about the same time as parallel work in Isadore Fankuchen\'s laboratory[77] at the then Polytechnic Institute of Brooklyn. Lipscomb began by studying compounds of nitrogen, oxygen, fluorine, and other substances that are solid only below liquid nitrogen temperatures, but other advantages eventually made low-temperatures a normal procedure. Keeping the crystal cold during data collection produces a less-blurry 3-D electron-density map because the atoms have less thermal motion. Crystals may yield good data in the x-ray beam longer because x-ray damage may be reduced during data collection and because the solvent may evaporate more slowly, which for example may be important for large biochemical molecules whose crystals often have a high percentage of water.
Other important compounds were studied by Lipscomb and his students. Among these are hydrazine,[78] nitric oxide,[79] metal-dithiolene complexes,[80] methyl ethylene phosphate,[64] mercury amides,[81] (NO)2,[82] crystalline hydrogen fluoride,[83] Roussin\'s black salt,[84] (PCF3)5,[85] complexes of cyclo-octatetraene with iron tricarbonyl,[86] and leurocristine (Vincristine),[87] which is used in several cancer therapies.
Positions, awards and honorsGuggenheim Fellow, 1954[88]Fellow of the American Academy of Arts and Sciences in 1960.[89]Member of United States National Academy of SciencesMember of the Faculty Advisory Board of MIT-Harvard Research JournalForeign Member of the Royal Netherlands Academy of Arts and Sciences (1976)[90]Nobel Prize in Chemistry (1976)Five books and published symposia are dedicated to Lipscomb.[6][91][92][93][94]
A complete list of Lipscomb\'s awards and honors is in his Curriculum Vitae.[95]
illiam N. Lipscomb Jr., a Harvard chemistry professor who won a Nobel Prize in 1976 for his research on the structure of molecules and on chemical bonding, died on Thursday in Cambridge, Mass. He was 91.
His death was announced by his son, James. Dr. Lipscomb was a Cambridge resident.
A protégé of the two-time Nobel laureate Linus C. Pauling, Dr. Lipscomb was a pioneering researcher whose work on the chemical structure of boranes — compounds of boron and hydrogen — continued Dr. Pauling’s work at the California Institute of Technology in the 1940s.
In terms of practical applications, boron compounds have shown some promise in radiation therapy for treating brain tumors. But mainly the work significantly advanced basic knowledge of the way atoms bond together.
As Dr. Lipscomb said: “For me, the creative process, first of all, requires a good nine hours of sleep a night. Second, it must not be pushed by the need to produce practical applications.”
Dr. Lipscomb’s research involved developing X-ray diffraction techniques, usually used as a tool in physics, that allowed him to map the connection of the atoms in an important but puzzling group of compounds called boron hydrides. The electronic structure turned out to be not simple linear molecules but rather complex three-dimensional objects. Dr. Lipscomb was able to explain these structures for the first time.
Like Dr. Pauling, who won the Nobel in chemistry in 1954 and the Nobel Peace Prize in 1962, Dr. Lipscomb was an admired teacher. He was on the faculty at Harvard from 1959 until his retirement in 1990. Three of his doctoral students went on to win Nobel Prizes in chemistry.
Thanks for reading The Times.Subscribe to The Times“He was always accessible; the door to his office was always open,” said Roald Hoffmann, professor emeritus of chemistry at Cornell University and one of Dr. Lipscomb’s Nobel protégés in 1981. He added, “He created a beautiful and coherent body of work on an important group of molecules that had eluded simple description.”ImageWilliam N. Lipscomb Jr.William N. Lipscomb Jr.Credit...Charles Krupa/Associated PressA man of multiple talents known for his wry sense of humor and his signature string tie, Dr. Lipscomb was a classical clarinetist who performed in chamber groups and had been principal clarinetist with the Pasadena Civic Orchestra and the Minneapolis Civic Orchestra.
He was not above dropping comical elements into his scientific publications. In one 1972 paper, he noted: “We admittedly made this observation with the benefit of hindsight. This science is known as retrospectroscopy.”
William Nunn Lipscomb Jr. was born on Dec. 9, 1919, in Cleveland. When he was a year old, his family moved to Lexington, Ky. One of his early doctoral students gave him his nickname, Colonel, for his Kentucky roots. In 1973, the Honorable Order of Kentucky Colonels made him a member.
Receiving the proverbial chemistry set as a birthday gift at age 11, Dr. Lipscomb became obsessed with science. He later recalled creating “evil smells” using hydrogen sulfide to drive his two sisters out of his room and nearly blowing up the house while concocting gunpowder for homemade fireworks.
Despite his prowess, Dr. Lipscomb told his son, James, that he received a C in high school chemistry. His grade, based on just the final exam, demanded that he memorize the first 10 elements of the periodic table, but Dr. Lipscomb could not be bothered by such mundane tasks. “I could just look it up,” he said. “So I didn’t do it.”
Dr. Lipscomb attended the University of Kentucky on a music scholarship but graduated with a degree in chemistry in 1941. He went on to work with Dr. Pauling and earned his doctorate in chemistry at Caltech in 1946. He joined Harvard after 13 years at the University of Minnesota. From 1982 to 1990 he was on the board of directors of Dow Chemical.
Besides his son, James, and a daughter, Dorothy, from his marriage to the former Mary Adele Sargent, Dr. Lipscomb is survived by his second wife, Jean Evans; their daughter, Jenna; three grandchildren; and four great-grandchildren.
In his later years, Dr. Lipscomb was a regular participant in the annual Ig Nobel Prize awards in Cambridge, sponsored by the Annals of Improbable Research. “We have a bunch of Nobel Prize winners who hand out our prizes,” said Marc Abrahams, the founder of the Ig Nobels. “Bill ended up as our grand star. He played clarinet at the beginning and end of each show, and he narrated funny videos we post on our Web site. He just had great timing.”
Dr. Lipscomb may also be the only Nobel laureate featured in a YouTube video offering instructions on how to tie a string tie.Rudolph Arthur Marcus (born July 21, 1923) is a Canadian-born chemist who received the 1992 Nobel Prize in Chemistry[2] \"for his contributions to the theory of electron transfer reactions in chemical systems\".[3] Marcus theory, named after him, provides a thermodynamic and kinetic framework for describing one electron outer-sphere electron transfer.[4][5][6] He is a professor at Caltech, Nanyang Technological University, Singapore and a member of the International Academy of Quantum Molecular Science.Contents1 Education and early life2 Career and research2.1 Marcus theory of electron transfer3 Honors and awards4 External links5 ReferencesEducation and early lifeMarcus was born in Montreal, Quebec, the son of Esther (born Cohen) and Myer Marcus. His interest in the sciences began at a young age. He excelled at mathematics at Baron Byng High School. He then studied at McGill University under Dr. Carl A. Winkler,[7] who had studied under Cyril Hinshelwood at the University of Oxford. At McGill, Marcus took more math courses than an average chemistry student, which would later aid him in creating his theory on electron transfer.[8]
He earned a B.Sc. in 1943 and a Ph.D. in 1946, both from McGill University.[9][10] In 1958, Marcus became a naturalized citizen of the United States.
Career and researchAfter graduating, in 1946, he first worked at the National Research Council (Canada) [11] followed by University of North Carolina, and Polytechnic Institute of Brooklyn. In 1952, at the University of North Carolina, he developed Rice–Ramsperger–Kassel–Marcus theory by combining RRK theory with transition state theory. In 1964, he taught at the University of Illinois.[12]
Marcus theory of electron transferElectron transfer is one of the simplest forms of a chemical reaction. It consists of one outer-sphere electron transfer between substances of the same atomic structure likewise to Marcus’s studies between bivalent and trivalent iron ions. Electron transfer may be one of the most basic forms of chemical reaction but without it life cannot exist. Electron transfer is used in all respiratory functions as well as photosynthesis. In the process of oxidizing food molecules, 2 hydrogen ions, 2 electrons, and an oxygen molecule react to make an exothermic reaction as well as H2O (water). Due to fact that electron transfer is such a broad, common, as well as essential reaction within nature, Marcus\'s theory has become vital within the field of chemistry.
2H+ + 2e− + 1/2 O2 → H2O + heat
A type of chemical reaction linked to his many studies of electron transfer would be the transfer of an electron between metal ions in different states of oxidation. An example of this type of chemical reaction would be one between a bivalent and a trivalent iron ion in an aqueous solution. In Marcus\'s time chemists were astonished at the slow rate in which this specific reaction took place. This attracted many chemists in the 1950s and is also what began Marcus\'s interests in electron transfer. Marcus made many studies based on the principles that were found within this chemical reaction, and through his studies was able to create his famous Marcus theory. This theory gave way to new experimental programs that contributed to all branches within chemistry.[13]
Honors and awardsHonorary degrees were conferred to Marcus by the University of Chicago in 1983, by the University of Goteborg in 1986, by the Polytechnic Institute of Brooklyn in 1987, by McGill in 1988, by Queen\'s University in 1993, by the University of New Brunswick also in 1993, by the University of Oxford in 1995, by the University of North Carolina at Chapel Hill in 1996, by the Yokohama National University in Japan also in 1996, by the University of Illinois at Urbana–Champaign in 1997, by the Technion – Israel Institute of Technology in 1998, by the Technical University of Valencia (Spain) in 1999, by the Northwestern University in 2000, by the University of Waterloo in 2002, by the Nanyang Technological University (Singapore) in 2010, by the Tumkur University (India) in 2012, by the University of Hyderabad (India) also in 2012, and by the University of Calgary in 2013. In addition, he was awarded an honorary doctorate from the University of Santiago, Chile in 2018.
Before receiving the Nobel Prize in 1992,[2] Marcus received the National Medal of Science in 1989,[14] the Irving Langmuir Award of the American Chemical Society in 1978,[15] the Willard Gibbs Award in 1988, the Theodore William Richards Award in 1990, the Pauling Medals in 1991, and the Remsen and Edgar Fahs Smith Awards in 1991, the Peter Debye Award of the American Chemical Society in 1988, the Robinson Award in 1982, the Centenary Medals of the Faraday Division of the Royal Society of Chemistry in 1988, Columbia University\'s Chandler Medal in 1983, Ohio State\'s William Lloyd Evans Award in 1990, the Wolf Prize in Chemistry in 1985 and the Hirschfelder Prize in Chemistry in 1993. Marcus has been a Member of the National Academy of Sciences since 1970, and a Member of the American Academy of Arts and Sciences since 1973.[15] He won the Wolf Prize in Chemistry in 1984.[16]
He also received a Professorial Fellowship at University College, Oxford from 1975 to 1976.
He was elected to the National Academy of Sciences in 1970, the American Academy of Arts and Sciences in 1973, the American Philosophical Society in 1990, received honorary membership in the Royal Society of Chemistry in 1991, and in the Royal Society of Canada in 1993.[17] He was elected a Foreign Member of the Royal Society (ForMemRS) in 1987.[1]


Buy Now

Related Items:

New 1976 American Freedom Train Enamel Red White & Blue Keychain 5/8

New 1976 American Freedom Train Enamel Red White & Blue Keychain 5/8" x 3/8"

$12.99



Vintage Silver 925 Gland England Birmingham 1976 Thimble James Swann & Son stamp picture

Vintage Silver 925 Gland England Birmingham 1976 Thimble James Swann & Son stamp

$300.00



The Legion of Monsters #1 (1975) VG/ FN 1st Manphibian, Curtis, Marvel Comics picture

The Legion of Monsters #1 (1975) VG/ FN 1st Manphibian, Curtis, Marvel Comics

$99.99



Powered by WordPress. Designed by WooThemes