Chemical & Chemical Engineering News (80th Anniversary Issue), Vol. 81, No. 36, 2003, Sept. Edited by X. Lu Introduction


GREGORY H. ROBINSON, THE UNIVERSITY OF GEORGIA



Yüklə 2,68 Mb.
səhifə36/44
tarix29.07.2018
ölçüsü2,68 Mb.
#59552
1   ...   32   33   34   35   36   37   38   39   ...   44

GREGORY H. ROBINSON, THE UNIVERSITY OF GEORGIA




Perhaps more than all others, the 13th element on the periodic table is one of utter contrasts. Although once highly valued as a "precious metal," the price of a kilogram of this element today is considerably less than one's morning cup of gourmet coffee; alloys of this element are often dense and durable, yet the pure element is a light and soft metal; although this element is reasonably reactive, it is also readily passivated, rendering it essentially rustproof.

Even its position on the periodic table--gracefully residing between the only nonmetallic element of group 13, boron, and the oddly mercurial metal of gallium--is perhaps indicative of how extraordinary element 13 is. There is not even complete agreement on the spelling and pronunciation: Americans typically employ aluminum, while significant portions of the remaining English-speaking world prefer aluminium. Aluminum, like most elements, has its share of interesting trivia: For example, a 2.73-kg pyramid of "precious" 1884 aluminum metal sits atop the Washington Monument.

While Hans Christian Oersted is acknowledged as the first to isolate aluminum in 1825 in Copenhagen, Denmark, the eminent German chemist Friedrich Wöhler is generally regarded as the first to secure a pure sample of the element by chemical reduction in 1827. The intriguing international tale of the discovery of the economical production of aluminum independently by two young men, the American Charles M. Hall and the Frenchman Paul L. T. Héroult, via electrolysis of alumina dissolved in cryolite, is well documented. Certainly, the aluminum industry as we know it today is due to the creative genius of Hall and Héroult. My fascination with aluminum, however, has less to do with the actual element and more with the relationship between Frank Fanning Jewett and Hall.

Jewett, educated at Yale University in chemistry and mineralogy, had a passion for travel. Indeed, he studied briefly at Universität Göttingen, spending time in the laboratory of Wöhler. In 1880, the 36-year-old Jewett was appointed professor of chemistry and mineralogy at Oberlin College. Thus, the stage was set for the well-traveled professor and the prodigious student. Jewett is the (most often) anonymous "professor" at Oberlin who opined to his chemistry class, where Hall was in attendance, that great financial rewards awaited the person who could devise an economical means to produce aluminum metal from its ubiquitous ore. The role of Jewett in Hall's life proved crucial in the seminal discovery that would ultimately spawn the Aluminum Co. of America, Alcoa (2002 revenues of $20.3 billion), and the worldwide aluminum industry.

The initiative and drive of Hall remain impressive. For an amateur scientist to doggedly pursue a scientific problem of such magnitude, and ultimately succeed in such an endeavor, is almost unimaginable. Records indicate that Jewett was a counselor, mentor, adviser, and friend to Hall. Furthermore, Jewett often provided materials and laboratory space to the budding entrepreneur. Jewett, a modest man by all accounts, apparently was not interested in sharing the praise, fame, or financial rewards that would soon befall his student. As Oberlin College's Norman C. Craig has so elegantly stated, Jewett was "content to report to his Yale classmates that his greatest discovery was the discovery of a man--Charles Hall" [Chem. Heritage, 15, 36 (1997)].

In my mind, the Jewett-Hall relationship epitomizes the idyllic professor-student dynamic. It is this relationship that I envision when I am working with students: a synergistic pursuit of the scientific unknown. To be sure, the stakes are much lower in my day-to-day struggles in the laboratory. The problems that my students and I face are much smaller in magnitude, and any potential immediate impact is often ambiguous.



Nonetheless, the Jewett-Hall relationship drives me in an oddly personal manner as I strive to improve my teaching skills and hone my research capabilities. Might my perspective on this relationship be a rather naive interpretation? Almost certainly. Is this simply an outdated commentary on the contemporary professor-student dynamic? Most clearly. Could this all be little more than a "nonprofessorial" waste of time? Absolutely not! The professor-student dynamic represents much of what I find uniquely attractive in academia. I have observed parallels to the Jewett-Hall relationship in athletics: In tennis, it is that perfectly executed service ace down the middle of the court; in basketball, it is that gracefully arching jumpshot from the corner, hitting "nothing but net"; in golf, it is that splendid tee shot on the par 5, 18th hole--you know, that one shot that keeps bringing you back time and time again.



Gregory H. Robinson is a distinguished research professor of chemistry at the University of Georgia. His research interests, the organometallic chemistry of the main-group metals, are tempered by his recent obsession with golf. The author acknowledges the gracious assistance of Norman C. Craig (Oberlin College) and Richard K. Hill (University of Georgia) with this essay.



ALUMINUM AT A GLANCE

Name: From Latin alumen,alum.

Atomic mass: 26.98.

History: Discovered in 1825 by Danish chemist Hans Christian Oersted.

Occurrence: Aluminum is the most abundant metal in Earth's crust, but it is not found free in nature. Today, nearly all of the world's aluminum is obtained by isolation from aluminum oxide derived from bauxite ore.

Appearance: Silvery white, lightweight metal.

Behavior: Soft, nonmagnetic, and nonsparking. Pure aluminum is easily formed, machined, and cast, and it can be alloyed with a variety of metals. It is also a good conductor of electricity and an excellent reflector of radiation. The metal is generally nontoxic but can be harmful when ingested.

Uses: Used to make cans, kegs, wrapping foil, and household utensils. It has numerous applications in the vehicle, aircraft, and construction industries.

GALLIUM

OLIVER SACKS, NEW YORK CITY




I don't really have a favorite element--I love them all. But the first that pops into my mind, at least today, is gallium.

Why gallium? Not a common element, not one likely to be lying around the house, but one I was introduced to quite early on, by my Uncle Tungsten (as we used to call my Uncle Dave, who owned a tungsten lightbulb factory). Uncle possessed what was, to my eyes, a quite wonderful thermometer that he used for testing the temperature in the furnaces in his factory. This, he showed me, contained gallium; it was the only stuff such a thermometer could contain, for gallium had the widest temperature range of any metal in the liquid state: It would melt in the heat of the hand but not boil until well over 2,000 °C (higher than the melting point of platinum or of the quartz of which the thermometer itself was made).






Sacks
Uncle gave me a lump of gallium to play with, and I can feel to this day the intense surprise I experienced when this lump, by no means soft, started to melt and trickle through my fingers as I held it. I later used a mold to make a teaspoon from it, and I would give this to unsuspecting guests and watch gleefully as they tried to stir their tea with it, only to find the seemingly solid spoon getting shorter and shorter and ending up as a glittering puddle at the bottom of the glass. And glitter it did--one had only to slosh liquid gallium round a hemispherical bowl to get an instant, brilliant, concave mirror.

A strange optical illusion appears if one melts gallium in a cup: There seems to be a transparent liquid skimming and floating above a silver background. Is this due to the strongly concave meniscus of liquid gallium? I do not think I ever saw this, by contrast, with mercury, which has a strongly convex meniscus. And once melted, gallium may remain liquid, superfluid, for many hours, even if the room temperature is well below its melting point. The liquid may form a skin, wrinkled with fine lines, and when it finally solidifies, it may do so in shallow quadrangular prisms and zigzags like medieval fortifications.

I also had a little stick of indium, another element that intrigues me, partly because, like tin and zinc, it emits a "cry" or "squeal" when bent. (With my little bar, it was more like a crackling.) One day, just recently, I carelessly left the indium on top of some gallium I had in a bowl, and I was startled to find, within hours, that the bar seemed to have partly dissolved and that there was now a pool of liquid metal at the bottom of the bowl, despite its being a rather cold day. Clearly, the two elements, merely by being in contact, had fused together to form a eutectic alloy with a substantially lower melting point than that of pure gallium. I was reminded of how Berzelius had been sent samples of metallic sodium and potassium, which were put together in the same container for convenience, and when he opened the package, he found only a pool of liquid metal, the two elements having spontaneously alloyed at room temperature, just as my indium and gallium had.

When I came to learn about the periodic table and its history, I was intrigued to learn that gallium was the first element to be predicted by Mendeleev based on its place in Group III (he called it "eka-aluminum"), and how this prediction was vindicated, just six years later, helping to convince Mendeleev's critics of the fundamental truth of his periodic law.



Many decades later, I was fascinated to learn that there was a huge pool containing 200 tons of ultrapure liquid gallium deep beneath the Caucasus, an essential part of the Soviet solar neutrino detector. A passionate swimmer, I had fantasies of swimming, or rather floating, on this unique lake of metal. I was shocked when I read, a few years ago, that thieves had come by with siphoning equipment one night and almost managed to steal the whole lot. The great gallium heist was foiled only at the last moment.



Oliver Sacks is a neurologist practicing in New York City. He is the author of "Uncle Tungsten: Memories of a Chemical Boyhood," as well as "Awakenings" and "The Man Who Mistook His Wife for a Hat."



GALLIUM AT A GLANCE

Name: From the Latin Gallia, an old name for France.

Atomic mass: 69.72.

History: Discovered spectroscopically in 1875 by French chemist Paul-Emile Lecoq de Boisbaudran. In the same year, he obtained the free metal by electrolysis of a solution of the hydroxide in KOH.

Occurrence: Gallium minerals are rare, but up to 1% gallium occurs in the ores diaspore, sphalerite, germanite, and bauxite. It is also recovered as a by-product of burning coal.

Appearance: Silvery metal. Extremely soft and can be cut with a knife. The metal expands upon solidifying.

Behavior: One of the few metals (with mercury, cesium, and rubidium) that can be a liquid at room temperature. It has one of the longest liquid ranges of any metal and has a low vapor pressure even at high temperatures. Gallium salts generally have low toxicity.

Uses: Gallium arsenide is capable of converting electricity directly into coherent light and is a key component of LEDs (light-emitting diodes) and some integrated circuits. Gallium is also used in semiconductors and solid-state devices, microwave equipment, low-melting alloys, mirrors, and high-temperature thermometers. Radioactive gallium is used in medical imaging.

INDIUM

CELIA HENRY, C&EN WASHINGTON




A couple of years ago, composer Richard Bone was approached to present a piece at an electronic music festival in St. Petersburg, Russia.

"I hadn't performed in 20 years, but for some unknown reason I said yes," he says. He was afraid that getting his equipment to St. Petersburg would be a "nightmare," so he decided to compose a "minimal piece of slow-moving music and use a video to accompany it."

Bone found inspiration for that piece in the most unlikely of places--the Periodic Table of the Elements. The combination of music and video is known as "Indium," a 30-minute composition consisting of three distinct sections. In the accompanying video, the scene changes every 10 minutes to match the changes in the music. It falls into the genre of experimental music known as ambient music.

Bone got his start in pop music and says that he is still something of a pop songwriter, because most of his pieces are only three or four minutes long. "This was the first time I had ever attempted to compose something that was 30 minutes long and try to keep it interesting and moving," he says. It was written as three separate pieces that Bone then had to flow into each other to make one continuous piece of music.

Interestingly, although indium ultimately provided the inspiration for the finished piece, the idea was originally planted by another element. A fan of the Spanish surrealist painter Joan Miró, Bone visited the Fundació Joan Miró in Barcelona, where he saw the "Mercury fountain," which was created by Alexander Calder for the 1937 World's Fair in Paris. In this sculpture, liquid mercury is piped from a pool and flows down a series of curved metal pieces back into the pool. The sculpture was created as a political statement against the seizure of the Almaden mercury mines by Francisco Franco during the Spanish Civil War.




MUSIC TO MY EARS Richard Bone's musical tribute to indium.
"I could watch that mercury flow forever," Bone says. "I came back and wanted to do a piece that slowly evolved like flowing mercury."

But he didn't want to call a piece "Mercury" because that had already been done, so he looked elsewhere for an appropriate name. "I somehow stumbled across a periodic table and saw the word indium," Bone says. "I'd never heard of it, so I looked up the description of it. The description almost perfectly described the music and the video that I wanted to do."

Indium is a soft, lustrous, silvery white metal. It is useful for making alloys with low melting points. For example, an alloy of 24% indium and 76% gallium is a liquid at room temperature. The element was discovered spectroscopically in 1863 by Ferdinand Reich and his assistant Hieronymous T. Richter while they were searching for thallium in zinc ore. The element was named for the brilliant indigo line in its spectrum. The pure metal was first isolated in 1867 by Richter. Until 1924, about a gram represented the entire world supply of the isolated metal. Indium was originally thought to be rare, but it is actually about as abundant as silver.

The definition of indium that Bone found "evoked in me the music I was creating." He printed the definition and taped it over the keyboard where he could see it as he worked on the piece. "There was something about the description of that element that seemed to capture what I wanted to create in the music," he says.

He took the definition to Jim Karpeichik, a local videographer, and told him that he wanted to create a visual that evolved and moved very slowly. Karpeichik filmed ocean scenes at the beach in Rhode Island, slowed them down, and colorized them.

Bone's approach to composing "Indium" was unusual for him. Typically, he improvises on the keyboard, looking for an unusual sound that he hasn't worked with before that can inspire him melodically. The work slowly starts to evolve from there. Often, he doesn't have a title for a work until he's done. "This was going to be a live performance piece, originally," Bone says. "I needed to have some very clear concept of what I was going to do."

The music festival fell through when the Russian government declined to fund it, but the festival producers also owned a record company. The Russian record company Electroshock Records released the album "Indium" in December 2002. In the U.S., it is exclusively available from the online ordering sites http://www.eurock.com and http://gemm.com. Another interesting tidbit, given that this is C&EN's 80th or "mercurial" anniversary: The second track on the album is called "Mercurial Wave."

Celia Henry is an associate editor for science, technology, and education at C&EN. A music lover with broad-ranging tastes, Celia discovered a new musical genre while writing this essay.


INDIUM AT A GLANCE

Name: From the Latin indicum, violet or indigo. It was identified by the bright violet light it emitted during spectroscopic analysis.

Atomic mass: 114.82.

History: Discovered in 1863 by German chemists Ferdinand Reich and Hieronymous T. Richter while looking for traces of thallium in samples of zinc ores.

Occurrence: Typically occurs along with zinc, iron, lead, and copper ores. Canada produces the majority of the world's supply of indium.

Appearance: Soft, silvery white metal with a brilliant luster.

Behavior: Moderately toxic by ingestion and affects the liver, heart, and kidneys. It is a suspected teratogen.

Uses: Alloyed with other metals to give them a lower melting point. Also used in transistors and thermistors and to wet glass.

THALLIUM

JEROME O. NRIAGU, UNIVERSITY OF MICHIGAN, ANN ARBOR




To many people, thallium is synonymous with rat poison. It is more toxic to mammals than mercury, cadmium, and lead and has been responsible for many deliberate, accidental, occupational, and therapeutic poisonings of people since its discovery in 1861. For the past 30 years or so, I have been fascinated by the risks associated with the ongoing release of this highly toxic and unpredictable metal into our environment.

Public angst and concern were first drawn in the late 1960s to reports of widespread contamination of the Great Lakes' ecosystems with toxic metals. At the time, I was a graduate student at the University of Toronto. Of the various groundbreaking studies on the hazards of heavy metals in the Great Lakes basin, I was captivated by reports that alluded to the fact that symptoms typical of thallium poisoning were being observed in many wildlife populations in the Great Lakes basin, and especially one report that claimed that nine out of the 34 bald eagles found sick or dying in 1971-72 in parts of the basin in the U.S. were poisoned by thallium. For some inexplicable reasons, these early reports on poisoning of wildlife by thallium were ignored by the scientific community and no efforts were made to conduct necessary follow-up studies. Nevertheless, they aroused my curiosity and served as my initiation into the literature on the environmental chemistry and toxicity of thallium.

Thallium is an enigmatic element to study because of its highly divergent properties predicated upon its oxidation status. Its paradoxical nature became obvious after its discovery as attempts were made to place the new element in the periodic classification of the elements. The physical properties of elemental thallium, including specific gravity, hardness, appearance, melting point, and electrical conductivity, were found to be similar to those of lead. The chemical properties of many Tl(I) (thallous) salts resemble those of lead, but the valency and other features of thallium were found to be too divergent from the lead family of elements.




CHECKUP Scan of a normal human heart after injection of radioactive thallium-201. The test can reveal defects in blood supply.
On the other hand, Tl(I) resembles the alkali metals in flame spectra; solubility of the hydroxide, sulfate, and carbonate in water; ready oxidation of the metal in air; existence of thallium alums; and isomorphism of some of its salts with those of potassium, cesium, and rubidium. But the absence of isomorphism and divergent properties of many of its common salts precluded thallium from the family of alkali metals. In his periodic classification of the elements, Dmitry Mendeleyev offered the clinching argument in favor of placing thallium among the aluminum group of elements based on the chemical properties of thallic salts. Variations in chemical personality and nonconforming behavior of thallium make for a fertile field for research but provide a minefield in any attempts to adequately assess its risks in the environment.

Thallium elicits some of the most complex and serious toxicities in living things, involving a wide range of organs and tissues. A recent study with planktonic communities showed that Tl3+ ions, a common form in aquatic environments, are about 34,000 times more toxic than cadmium ions [Environ. Sci. Technol., 37, 2720 (2003)]. This is serious, considering that organisms (especially mammals) at the top of the food chain are the most susceptible to thallium toxicosis. Recently, we have found elevated levels of thallium in Great Lakes fish [Bull. Environ. Contam. Toxicol., 67, 921 (2001)], which has raised the issue of bioaccumulation and biomagnification of this element in aquatic food webs and the potential risks to fish-eating members of the food chain. I keep wondering whether the dismissal of the early reports of thallium poisoning of bald eagles and other wildlife should, in fact, be regarded as a major scientific oversight of our time.

Compared with many heavy metals, thallium has a short history and what appears to be a rosy future. Its traditional uses (in rodenticides and insecticides, pigments, wood preservatives, and ore separation; in mercury lamps to increase the intensity and spectrum of the light; as catalysts in chemical syntheses; and so forth) are being phased out in deference to its toxicity. At the same time, there is increasing demand for thallium in the high-technology and future-technology fields.

Among the growing uses for thallium are in the semiconductor and laser industry, in fiber (optical) glass, in scintillographic imaging, in superconductivity, and as a molecular probe to emulate the biological function of alkali-metal ions. Recycling of thallium in commercial products is not yet a serious business, and one has to be concerned about the long-term environmental effects of the growing technological applications of one of the most toxic and eccentric metals known to humankind.



Jerome O. Nriagu is a professor in the School of Public Health at the University of Michigan, Ann Arbor. He is the editor of "Thallium in the Environment" as well as many other volumes on heavy metals in the environment.


THALLIUM AT A GLANCE

Name: From the Greek thallos, green twig or shoot, after the color of its spectral lines.

Atomic mass: 204.38.

History: Discovered in 1861 by Sir William Crookes.

Occurrence: Thallium can be obtained as a by-product of producing sulfuric acid or refining zinc or lead.

Appearance: Silvery white metal.

Behavior: Thallium compounds are extremely

Uses: TI is used in fiber and low-melting glass. Many traditional uses have been phased out because of its toxicity.


CARBON

GEORGE A. OLAH, UNIVERSITY OF SOUTHERN CALIFORNIA




Carbon is the key element not only of terrestrial life but also of minerals (carbonates) and fossil fuels (oil, gas, and coal), and it is a minor but essential component of our atmosphere. Carbon is produced in the stars by nuclear synthesis from hydrogen that originated from the initial Big Bang. Over the eons, asteroids hitting Earth may have carried carbon to our planet.

Elemental carbon is found in nature as its allotropes--diamond and graphite--which are of vastly differing abundance and thus also of differing value. In the 1980s, a new group of allotropes called fullerenes or buckyballs--named after R. Buckminster Fuller, who designed famous geodesic domes resembling soccer balls--was recognized, first spectroscopically (for which Robert F. Curl Jr., Sir Harold W. Kroto, and Richard E. Smalley received the 1996 Nobel Prize in Chemistry) and later produced in electric discharge devices using carbon electrodes (Donald R. Huffman and Wolfgang Krätschmer). These new carbon allotropes promise significant applications.

Carbon has a remarkable ability to bind with itself to form chains, rings, and complex structures. The variety of carbon compounds with bound hydrogen (hydrocarbons) and other elements (oxygen, nitrogen, and phosphorus, for example), which are generally called organic compounds, is practically unlimited.




C60
The very heart and essence of chemical transformations and hence of modern chemistry is an understanding of chemical reactions and their intermediates. Carbocations (the positive ions of carbon compounds) and related electron-deficient species represent the most important intermediates in all of chemistry. I was fortunate to have discovered ways to observe carbocations as persistent, long-lived species.

The related development and use of superacids, billions of times stronger than sulfuric acid, and superelectrophiles have changed the field of chemistry. The realization that carbon, in many of its electron-deficient species, including carbocations, can simultaneously bind (coordinate) to five, six, and even seven neighboring atoms or groups is significant. This has extended Kekulé's concept of the limiting four valency of carbon to higher coordinate carbon compounds and has opened up new fields of hypercarbon chemistry.

My discoveries also provided insights into the electrophilic activation of C–H and C–C single bonds and formed the basis of the development of new and improved hydrocarbon transformations. These have significant applications in the petroleum and chemical industries for improved production of high-octane gasoline (via alkylation and isomerization) and the direct conversion of methane (natural gas) to methanol and higher hydrocarbons without producing syngas.

From plant life over the ages, new fossil fuels can be formed. The process is so slow, however, that within our human life span we do not have time for nature to replenish what we are rapidly using up. A challenging new approach that we are pursuing is to reverse the process and produce hydrocarbons from carbon dioxide and water via methanol, thus chemically recycling carbon dioxide. In the laboratory, we already know how to do this, and progress is being made toward bringing about the feasibility of such an approach. The limiting factor is the energy needed for generating hydrogen from water. Using alternative energy sources--but first of all atomic energy, albeit improved and made safer--will eventually give us needed energy.

Much is said these days about a "hydrogen economy," emphasizing hydrogen as the clean, inexhaustible fuel of the future. However, the safe handling and dispensing of volatile hydrogen--for which no infrastructure exists--is difficult and costly.

I believe a much preferable way of storing hydrogen is in the form of methyl alcohol ("methanol economy"). Methanol is a convenient liquid that can be produced by reduction of carbon dioxide in the atmosphere. It can be catalytically converted into ethylene and propylene and through them to higher hydrocarbons. This can provide an inexhaustible source of hydrocarbon products and fuels, which are now obtained from oil and gas. Furthermore, in recent years, with colleagues at California Institute of Technology and the Jet Propulsion Laboratory, we have also developed a new, direct methanol fuel cell that produces electric power without the need of hydrogen. Thus, methanol is both a fuel and a source of hydrocarbons. By recycling excess CO2 into methanol instead of just storing or sequestering it, we can also mitigate global warming. It is to this effect that a major research effort, with my colleagues associated with the Loker Hydrocarbon Research Institute at the University of Southern California, is directed.





George A. Olah is a professor of chemistry and director of the Loker Hydrocarbon Research Institute at the University of Southern California. He won the 1994 Nobel Prize in Chemistry for his work on carbocation and hydrocarbon chemistry.



CARBON AT A GLANCE

Name: From the Latin carbo, coal.

Atomic mass: 12.01.

History: Known since ancient times.

Occurrence: Carbon occurs naturally as crystalline graphite or diamond, or amorphously in charcoal, carbon black, coke, and white carbon. A new molecular allotrope, C60, or buckminsterfullerene, was discovered in 1985.

Appearance: Nonmetal solid. Graphite is black or silver-black. Diamond comes in various colors.

Behavior: Not volatile. Readily bonds with many elements and itself, forming millions of compounds.

Uses: Essential component of all organic molecules, including nucleic acids, proteins, and carbohydrates. The C-14 isotope is used for carbon dating. Graphite is used as a lubricant and in pencils. In addition to being valued as a gem, diamond is often used as an abrasive. Carbon dioxide has numerous uses, including refrigeration.

SILICON

ROALD HOFFMANN, CORNELL UNIVERSITY




Similarity and difference nourish the mind at play, and not just for a boy whose fate was to be a refugee. My introduction to the periodic table actually came through a science story about silicon buried in the pages of a comic book I found lying around on the S.S. Ernie Pyle, the troop carrier that brought us to America in 1949.

The story was into similarity; it invoked no less an icon of analogical thinking than Dmitry Ivanovich Mendeleyev. Silicon was under carbon; surely it could not be denied its Mendeleyevian birthright, to give us all that carbon did? In the heyday of silicon chemistry, it was just a matter of time and chemistry. Or maybe of different ambient conditions; silicon-based life on another planet made such good science fiction sense.

I was anxious to learn English, to speak without an accent, to fit in; of the life-enhancing differences, such as that between boys and girls, there were only stirrings. I lapped up that vision of an alternative, yet like, universe.

But silicon is different. Yes, one can make Si analogs of hexane and cyclohexane. The differences emerge earlier in reactivity than in structure. Most importantly, unsaturation is strongly disfavored for silicon, as it is for other main-group elements below the first row. At a normal Si–Si distance, p bonding is just not worth very much. So double- and triple-bonded molecules have high-lying filled and low-lying unfilled orbitals. This makes for reactivity to acids, bases, and radicals. Silylenes, SiR2, are also very stable; put these trends together and an equilibrium such as H2Si = SiH2 H3Si – SiH, highly endothermic for carbon, is nearly thermoneutral and faces a small activation energy for silicon. Or, should you want something still more striking, it takes less energy to dissociate H2Si = SiH2 into two silylenes than it takes to break the single bond in H3Si–SiH3. Five and six-coordinate structures also become accessible for neutral compounds of Si.

Just how different silicon is shows up in the equilibrium structure of Si2H2. In a triumph of computational chemistry (not by me, alas), Si2H2 was predicted to have the striking nonclassical dibridged structure, which was then confirmed experimentally.

The bond energies of C–E and Si–E (E = an element) are similar (within 50 kJ per mol) for a large variety of Es. The exception is Si–O (and Si–F), which is nearly 100 kJ per mol stronger than C–O. Coupled to the unhappiness of Si with unsaturation, silicon's love of oxygen leads to an overwhelming difference between a world of CO multiple bonds and the universe of silicates. Contrast gaseous CO2 and solid polymeric SiO2.

Poor carbon, to have no 2- and few 3-D polymers. Compare clays, micas, zeolites, aerogels, amethyst and carnelian, lapis lazuli, asbestos, porcelain, and glass--just to pick a sampler of silicates.

Silicon can certainly support chirality. Its chemistry sports helices. So why has this element essentially been dealt out of life, at least out of animal biochemistry? The abundance of silicon in Earth's crust is second only to oxygen. Much less common elements--take copper or molybdenum--are co-opted by life; one would have thought that nature's evolutionary tinkering surely would have found ample use for silicon.

I'm not quite fair. The element, through silica, is of immense utility as a structural material in diatoms and radiolarians. A typical one, Cyclotella cryptica, is 22% SiO2 by dry weight. There are a lot of those little guys, too. Silicates are also essential for higher plants. Horsetails, which once grew into tall trees, accumulate enough silica to make them used as scouring brushes. There's a lot of silica in rice and in grasses in general.

But almost none in bacteria and animals. It's a great puzzle. Though there are advocates of the way of clay, it appears that the road to life was through isomeric variability and a balancing act of metastable yet kinetically persistent small organic molecules and one-dimensional polymers. Poor silicon, capable of supporting isomerism but pulled away from Si–Si catenation by that seductive bond to abundant oxygen!

How unexpected then (little help from science fiction here) is our here-and-now-and-still-growing in silico world. This is silicon's revenge! Or the evolution of electronics and computers might be seen as only the latest extension of human cultural appropriation of those elements underused in biology. Following (for silicon) that constant of civilization--ceramics. And glass.

In the excitement of building a theory, in the desperate and thrilling labor of trying to understand, the primal urge is to simplify. Not just in science, though our craft is prone to special reductive tendencies. Simplifying to excess, we make things the same. But the beauty of this world is as much in the workings of chance and in the infinite variety that being not quite the same provides. So it is for silicon and carbon, as it is for people.



Roald Hoffmann is the Frank H. T. Rhodes Professor of Humane Letters at Cornell University. He received the Nobel Prize in Chemistry in 1981.


SILICON AT A GLANCE

























Name: From the Latin silex, flint.

Atomic mass: 28.09.

History: Jöns J. Berzelius is credited with discovering silicon in 1824.

Occurrence: On Earth, silicon is the second most abundant element, making up 25.7% of the crust by weight.

Appearance: Amorphous silicon is a brown powder; crystalline silicon is dark gray with a bluish tinge.

Behavior: Pure Si is covered in a layer of SiO2, rendering it inert to air or water.

Uses: Si is used in lasers, transistors, and other solid-state devices.


GERMANIUM

BETHANY HALFORD, C&EN WASHINGTON


If the elements in the periodic table were pictures in a high school yearbook, germanium would be that geeky, nondescript kid that no one remembers. Part of germanium's problem is its status as the middle child of the main group, sandwiched between elemental overachievers carbon and silicon on one side and tin and lead on the other.

When Mendeleyev conceived the periodic table, germanium hadn't even been discovered. I imagine him as a father with an overwhelming brood, leaving a place for an errant child, knowing that it has to be around somewhere. It doesn't even get respect from spellcheck, which insists on renaming it as a flower.

But as elements go, germanium is a late bloomer. It took the inventions of the transistor and crystal diode in the 1940s to transform it from chemical curiosity to important industrial material. Before 1945, only a few hundred pounds of the element were produced each year. But by the end of the 1950s, annual worldwide production had reached 40 metric tons.

Finding steady work in solid-state electronics, germanium held onto its high-tech reputation through the 1970s. And when sister silicon edged in on its territory, germanium moved on to more intriguing applications: polymerization catalysts, fiber-optic communication networks, and infrared night-vision systems.

The U.S. government even designated germanium as a strategic and critical material, calling for a 146,000-kg supply in the national defense stockpile for 1987. Like the brainy, adolescent wallflower who grows up to make a fortune as president of a high-tech company, germanium had made a metamorphosis.

Its cachet has come at a cost, though. In 1998, the price of 1 lb of germanium was almost $800; 1 lb each of silicon, tin, and lead could be bought for less than $5.

My own personal experience with germanium, and its high cost, came the summer before I started graduate school when I worked with noted main-group organic chemist Peter P. Gaspar at Washington University in St. Louis. I was to make several hundred grams of (CH3)2GeCl2 using Gaspar's clever synthetic protocol [Synth. React. Inorg. Met.-Org. Chem., 20, 77 (1990)]. Perhaps because he was reluctant to let an inexperienced student like me work with pricey germanium powder--I had given him a purchase order for several thousand dollars' worth of the stuff--Gaspar suggested I call some chemical suppliers and see if any would be willing to discount their older stock. A little bit cocky, I was hurt at his lack of faith in my synthetic skills.

One supplier gave me a bargain price on some decade-old "five nines" pure material that had been taking up space on his shelves. Gaspar had said that less pure material tended to work better in the reaction, so I agreed. I figured the stoichiometry in my notebook for the germanium I needed if it was only 5/9 pure. I told Gaspar of the deal.

"I thought I told you that less pure material worked better," he said.

"Yeah, it's only five nines pure," I countered.

He stared at me, unmistakably annoyed. Irritated at his lack of enthusiasm, I skulked away. Only when the dusty old bottle arrived did I learn that "five nines" pure means 99.999%.

Gaspar once showed me a half-kilo ingot of germanium metal that a company had given him in the halcyon days before the government began stockpiling the stuff. I can still remember the sleek, cool feel of the dark metal. By the way Gaspar cradled the bread-loaf-shaped object in his hand, I could tell it was one of his treasures. I was paralyzed by the thought that he might actually ask me to grind the material up and use it.

For this essay, I called Gaspar to ask if he still had the ingot or if some unlucky student had to grind it up in the six years that had passed. "I keep it as a curiosity," he said, but conscious of its value, would not elaborate. "It's in a safe place."

When I asked Gaspar what he thinks about my theory of germanium as the main group's middle child, he said he could see the parallel, but added--in his wise professorial tone of voice--that germanium has never had plain-Jane status in chemistry. On the contrary, "the most important thing about germanium was the fact that it started as a twinkle in Mendeleyev's eye," he said. "It might have been found much later than it was if it weren't for the fact that it had to be there if the periodic table was correct."



Bethany Halford is an assistant editor at C&EN. Five nines of the time spent on this essay went into writing this sentence.



GERMANIUM AT A GLANCE

Name: From the Latin Germania, Germany.

Atomic mass: 72.64

History: In 1871, Mendeleyev used his periodic table to predict the existence of a silicon-like element. German chemist Clemens Winkler discovered it 15 years later.

Occurrence: are. Found in coal and compounds such as argyrodite.

Appearance: Gray, solid metalloid.

Behavior: Germanium is a semiconductor but exhibits poor conductivity.

Uses: Doped germanium materials are used to make transistors for miniature electronics.

TIN

SANDY GERRARD, ENGLISH HERITAGE




It was in 1979, while taking refuge from yet another blizzard in a rather inadequate shelter on top of a Welsh mountain, that the opportunity of digging in a somewhat more exotic location was first mentioned. At the time, I was an undergraduate studying archaeology at Lampeter in mid-Wales, and the new excavation opportunity was a tin mill at Colliford, somewhere in Cornwall. At this time, no British tin mill had ever been properly excavated, but I think it was the thought of six weeks in sunny Cornwall that persuaded me that it would be a good idea to be involved.

I will remember the first three days of that dig for the rest of my life. Dense fog, combined with torrential rain and powerful winds tragically known as the Fastnet Storm, coincided with our arrival. Once the winds had abated, the sun shone, and very rapidly it became apparent as we dissected the mill and the surrounding dressing floors that we were revealing a very complicated and informative story.

The mill, throughout its life, from at least 1507 until around 1600, had crushed tin from the nearby opencast quarry, but it had been abandoned and refurbished on several occasions, each time becoming more efficient. A wide range of artifacts provided an insight into the character of the tin processing operations, and domestic rubbish gave us a glimpse into the lives of the tinners who had worked here. The true complexity of the site only became apparent during the post-excavation process and preparation of the final report.




DIGGING DEEP View of the interior of the stamping mill at Colliford. The wheelpit is the water-filled, stone-lined channel on the right. The channel leading under the two near ranging rods carried material in suspension from the stamps situated adjacent to the wheelpit.
Understanding the mill proved challenging and rewarding, and it was during this process that I was smitten and decided I wanted to find out much more about the industry in which this mill had played a part. The landscape around the mill was littered with the earthworks and other structures left by the tinners. The obvious next step was to record and hopefully understand what could only be described as a confused mass of humps and bumps.

So it was on the last day of January 1983 that I set off with a plane table to tackle the earthworks in the valley bottom next to the mill. I really had no idea whether it would be possible or even worthwhile, as nobody else had ever tried to tackle this type of survey. I certainly did not know when I plotted the first point on the table that the results would be so informative that for the next 20 years I would be spending large blocks of time surveying and interpreting tinwork earthworks all over Cornwall and Devon.

The survey work revealed that it was possible to demonstrate exactly how the tinners had used different methods to extract tin. By doing a detailed analysis of the streamwork plans in particular, I could identify the precise methods used to extract the cassiterite and to recognize earthworks of different dates. Much has yet to be achieved regarding the absolute dating of the tinworks, but pollen analysis near the Colliford tin mill indicated that at least part of the tin streamwork was abandoned before 600 to 700 A.D. Many of the surviving streamworks in the southwest of England will be much more recent than this, as most probably belong to the late medieval period (1300 to 1500).

The scrutiny of streamwork earthworks has been the most productive aspect of the detailed survey of tinworking remains, but other types of tinwork lend themselves to this form of examination to a greater or lesser extent. Analyses of the surface workings associated with early forms of mining--including the shallow shafts known as lode-back pits and the opencast quarries known as openworks or beams--have provided a valuable insight into the earliest forms of mining. Together with investigations of the contemporary documentation, the surveys have allowed us to build a remarkable picture of the industrial, technological, and social character of early tin exploitation in Britain.



Sandy Gerrard is a designation archaeologist for English Heritage. He has directed several archaeological excavations and surveys and has published extensively on the early tin industry.



TIN AT A GLANCE

Name: From the Anglo-Saxon tin, named for the Etruscan god Tinia. The symbol is from the Latin word for tin, stannum.

Atomic mass: 118.71.

History: Known to ancient civilizations.

Occurrence: Often found as tin oxide, also called cassiterite. The metal can be isolated by heating the ore in the presence of carbon.

Appearance: Silvery white metal. Two known allotropes are white tin, which is metallic and malleable, and gray tin, which is brittle and powdery.

Behavior: Resists corrosion. Trialkyl- and triaryltin compounds are toxic.

Uses: Often used as a protective coating on other metals. Tin is alloyed with various metals to make foil, cans, solder, and pewter. Because it melts at a fairly low temperature, tin is ideal for casting. Bronze, an alloy of 80% copper and 20% tin, was used in ancient times to make tools and decorations. Stannic oxide and stannic chloride are used to make ceramic glazes and fabric treatments. Tin is also a major player in the Pilkington process for making glass panes.

LEAD

BASSAM Z. SHAKHASHIRI, UNIVERSITY OF WISCONSIN, MADISON




As a child in my native Lebanon, I collected ancient coins, digging Phoenician and Roman coins from olive groves near my hometown and from other locations. Several times, my parents took me to areas where archaeologists were carefully digging for artifacts. Occasionally, I would show some of my coins to one of the archaeologists and he would tell me in French or English what it was made of and how old it was. I was intrigued by the solidity of the coins, and I wanted to know more about their composition. I was told that most were alloys of copper. None had gold, but one or two may have had some silver. The coins felt heavy, and I wondered if there was lead in them.

One of my classmates was overweight, and we nicknamed him "lead." Later on, when we learned about the symbols of the elements, we referred to him simply as Pb, an abbreviation of the Latin plumbum, from which we also get the word plumbing. A Jesuit archaeologist told me that lead is mentioned in the Bible many times, that alchemists believed it was the oldest metal, and that they tried to change it into gold. He also told me that lead is poisonous. This fascinated me because I wondered how a solid piece of metal could be eaten. The Romans used lead to make pipes, some of which may still be in use. Later, I read that the Romans also used lead in cookware and drinking cups and that lead poisoning may have contributed to the fall of the Empire.

In high school and later as a freshman at the American University of Beirut, I learned more about lead. I was astonished to discover that lead is the end product of three series of naturally occurring radioactive elements and that lead has more than two dozen radioactive isotopes. In my sophomore year at Boston University, I learned that lead was in a gasoline additive--an organometallic compound called tetraethyl lead. I learned that lead is a major component of automobile batteries, that it was widely used in paint, and that it is used in soldering. Later, I learned that even in ancient times, some physicians believed that lead was poisonous, but it continued to be used in medicines and cosmetics until the 20th century.

Concerns over lead exposure in recent decades reflect heightened societal concerns for health and safety. All developed countries have banned two uses of lead that were once almost universal--tetraethyl lead and lead-based paint for residential use. The use of lead shot for hunting waterfowl has also been banned in the U.S. because bottom-feeding birds can ingest spent shot. Hunters must use steel shot instead. Older U.S. cities that have lead pipes or heavily soldered pipes are replacing them at considerable expense. There is continuing concern about lead glazes still used in some ceramics made in developing countries, but the lead found in fine china and crystal glass does not readily leach out and is not a concern.

Until recently, lead poisoning was diagnosed by its symptoms, but it is now diagnosed by analyzing its presence in blood by atomic absorption methods. The mechanism is not well understood, but lead is believed to bind to proteins involved in neurological signaling and development that otherwise would bind to calcium or zinc. Removing lead from the body safely remains a major challenge.




GETTING THE LEAD OUT University of Wisconsin, Madison, mascot Bucky Badger, named after Wisconsin's former lead miners, pours colorless potassium iodide solution into a colorless solution of lead nitrate, forming brilliant yellow, solid lead iodide.
Today, lead is essential to the manufacture and operation of many high-tech products. Lead solder is a reliable method of connecting transistors and other electronic components, and without leaded glass, we could not safely sit in front of our computer screens. Lead is the best material for nuclear radiation shielding and allows for the safe operation of CAT scans and other imaging diagnostics.

I have been a Wisconsin Badger for more than 30 years. The nickname of the state, and the mascot of the University of Wisconsin, Madison, are derived from lead. Lead mining in southwestern Wisconsin was one of the state's first industries. During the 1830s, lead miners, mostly young, single men, came up the Mississippi and Wisconsin Rivers from St. Louis and points south to spend their summers digging for lead. Not wanting to waste time on frills like housing, they lived in dugouts in the sides of hills. Some residents said they lived like badgers, a derogatory label they adopted with pride.



Bassam Z. Shakhashiri is professor of chemistry at the University of Wisconsin, Madison. His annual "Once Upon a Christmas Cheery in the Lab of Shakhashiri" is in its 34th year and is presented on PBS and cable stations.



LEAD AT A GLANCE

Name: From the Anglo-Saxon laedan, lead. The symbol is from the Latin word for lead, plumbum.

Atomic mass: 207.2.

History: Known to ancient civilizations. Alchemists believed lead to be the oldest metal and associated it with the planet Saturn. They also believed it could be transmuted into gold.

Occurrence: Usually obtained from galena (lead sulfide). Elemental lead is found only sparingly.

Appearance: Bluish white, soft metal.

Behavior: Lead is moderately toxic by ingestion and is a cumulative poison. Lead is very soft, highly malleable and ductile, a poor conductor of electricity, and very resistant to corrosion.

Uses: Used in batteries, glass, solder, radiation shielding around X-ray equipment and nuclear reactors, cable covering, plumbing, and ammunition. Lead was once used extensively in paints but has been phased out of most to eliminate health hazards.

NITROGEN

PETER NAGLER, DEGUSSA AG



Nitrogen is all around us, making up 78% of the air we breathe, yet we do not notice its presence; an atmosphere of 100% nitrogen, although nontoxic, is fatal. Leguminous bacteria, living in the root nodules of plants such as clover, have an advantage over other living things, as they can convert atmospheric nitrogen to nitrate, providing this essential element for the growth of legumes. This feat of biology was not matched by chemistry until 1914, when Fritz Haber established an industrial process for the manufacture of ammonia from atmospheric nitrogen, for which he received the Nobel Prize in Chemistry in 1918.

Although nitrogen was initially thought to be unreactive, this simple diatomic molecule does have interesting chemical properties. Some elements can "burn" in nitrogen, among them magnesium at 300 °C and lithium even at room temperature, producing crystalline metal nitrides. Complexes of molecular nitrogen with transition metals had been predicted for many years, but the first organometallic compound of dinitrogen, [Ru(NH3)5N2]Cl2, was only reported in 1965. Most of the fascinating chemistry of nitrogen is, however, reserved for its inorganic and especially organic compounds, notably amines, nitro compounds, and their more complex relatives.






FREEZE FRAME Nitrogen helps form the backbone of proteins. Shown here, a crystal structure of BRCA2 bound to single-stranded DNA.

Yüklə 2,68 Mb.

Dostları ilə paylaş:
1   ...   32   33   34   35   36   37   38   39   ...   44




Verilənlər bazası müəlliflik hüququ ilə müdafiə olunur ©genderi.org 2024
rəhbərliyinə müraciət

    Ana səhifə