Jump to content

Aluminium

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Ddurant (talk | contribs) at 05:44, 16 February 2009 (→‎See also: Referece to reducing aluminium toxicity). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Aluminium, 13Al
Aluminium
Pronunciation
Alternative nameAluminum (U.S., Canada)
AppearanceSilvery gray metallic
Standard atomic weight Ar°(Al)
Aluminium in the periodic table
Hydrogen Helium
Lithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine Neon
Sodium Magnesium Aluminium Silicon Phosphorus Sulfur Chlorine Argon
Potassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine Xenon
Caesium Barium Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury (element) Thallium Lead Bismuth Polonium Astatine Radon
Francium Radium Actinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium Rutherfordium Dubnium Seaborgium Bohrium Hassium Meitnerium Darmstadtium Roentgenium Copernicium Nihonium Flerovium Moscovium Livermorium Tennessine Oganesson
B

Al

Ga
magnesiumaluminiumsilicon
Atomic number (Z)13
Groupgroup 13 (boron group)
Periodperiod 3
Block  p-block
Electron configuration[Ne] 3s2 3p1
Electrons per shell2, 8, 3
Physical properties
Phase at STPsolid
Melting point933.47 K ​(660.32 °C, ​1220.58 °F)
Boiling point2743 K ​(2470 °C, ​4478 °F)
Density (at 20° C)2.699 g/cm3[4]
when liquid (at m.p.)2.375 g/cm3
Heat of fusion10.71 kJ/mol
Heat of vaporization284 kJ/mol
Molar heat capacity24.20 J/(mol·K)
Vapor pressure
P (Pa) 1 10 100 1 k 10 k 100 k
at T (K) 1482 1632 1817 2054 2364 2790
Atomic properties
Oxidation states−2, −1, 0,[5] +1,[6] +2,[7] +3 (an amphoteric oxide)
ElectronegativityPauling scale: 1.61
Ionization energies
  • 1st: 577.5 kJ/mol
  • 2nd: 1816.7 kJ/mol
  • 3rd: 2744.8 kJ/mol
  • (more)
Atomic radiusempirical: 143 pm
Covalent radius121±4 pm
Van der Waals radius184 pm
Color lines in a spectral range
Spectral lines of aluminium
Other properties
Natural occurrenceprimordial
Crystal structureface-centered cubic (fcc) (cF4)
Lattice constant
Face-centered cubic crystal structure for aluminium
a = 404.93 pm (at 20 °C)[4]
Thermal expansion22.87×10−6/K (at 20 °C)[4]
Thermal conductivity237 W/(m⋅K)
Electrical resistivity26.5 nΩ⋅m (at 20 °C)
Magnetic orderingparamagnetic[8]
Molar magnetic susceptibility+16.5×10−6 cm3/mol
Young's modulus70 GPa
Shear modulus26 GPa
Bulk modulus76 GPa
Speed of sound thin rod(rolled) 5000 m/s (at r.t.)
Poisson ratio0.35
Mohs hardness2.75
Vickers hardness160–350 MPa
Brinell hardness160–550 MPa
CAS Number7429-90-5
History
Namingfrom alumine, obsolete name for alumina
PredictionAntoine Lavoisier (1782)
DiscoveryHans Christian Ørsted (1824)
Named byHumphry Davy (1812[a])
Isotopes of aluminium
Main isotopes[9] Decay
abun­dance half-life (t1/2) mode pro­duct
26Al trace 7.17×105 y β+84% 26Mg
ε[10]16% 26Mg
γ
27Al 100% stable
 Category: Aluminium
| references

Aluminium (ˌæljʊˈmɪniəm, /ˌæljəˈmɪniəm/) or aluminum (/əˈluːmɪnəm/, see spelling below) is a silvery white and ductile member of the boron group of chemical elements. It has the symbol Al; its atomic number is 13. It is not soluble in water under normal circumstances. Aluminium is the most abundant metal in the Earth's crust, and the third most abundant element therein, after oxygen and silicon. It makes up about 8% by weight of the Earth’s solid surface. Aluminium is too reactive chemically to occur in nature as the free metal. Instead, it is found combined in over 270 different minerals.[11] The chief source of aluminium is bauxite ore.

Aluminium is remarkable for its ability to resist corrosion (due to the phenomenon of passivation) and its low density. Structural components made from aluminium and its alloys are vital to the aerospace industry and very important in other areas of transportation and building. Its reactive nature makes it useful as a catalyst or additive in chemical mixtures, including being used in ammonium nitrate explosives to enhance blast power.

Characteristics

Aluminium is a soft, durable, lightweight, malleable metal with appearance ranging from silvery to dull grey, depending on the surface roughness. Aluminium is nonmagnetic and nonsparking. It is also insoluble in alcohol, though it can be soluble in water in certain forms. The yield strength of pure aluminium is 7–11 MPa, while aluminium alloys have yield strengths ranging from 200 MPa to 600 MPa.[12] Aluminium has about one-third the density and stiffness of steel. It is ductile, and easily machined, cast, and extruded.

Corrosion resistance can be excellent due to a thin surface layer of aluminium oxide that forms when the metal is exposed to air, effectively preventing further oxidation. The strongest aluminium alloys are less corrosion resistant due to galvanic reactions with alloyed copper.[12] This corrosion resistance is also often greatly reduced when many aqueous salts are present however, particularly in the presence of dissimilar metals.

Aluminium atoms are arranged in a face-centered cubic (FCC) structure, which may explain its high melting point. Aluminium has a high stacking-fault energy of approximately 200 mJ/m².[13]

Aluminium is one of the few metals that retain full silvery reflectance in finely powdered form, making it an important component of silver paints. Aluminium mirror finish has the highest reflectance of any metal in the 200–400 nm (UV) and the 3000–10000 nm (far IR) regions, while in the 400–700 nm visible range it is slightly outdone by tin and silver and in the 700–3000 (near IR) by silver, gold, and copper.[citation needed]

Aluminium is a good thermal and electrical conductor, by weight better than copper. Aluminium is capable of being a superconductor, with a superconducting critical temperature of 1.2 kelvin and a critical magnetic field of about 100 gauss.[14]

Isotopes

Aluminium has nine isotopes, whose mass numbers range from 23 to 30. Only 27Al (stable isotope) and 26Al (radioactive isotope, t1/2 = 7.2 × 105 y) occur naturally; however, 27Al has a natural abundance of 99.9+ %. 26Al is produced from argon in the atmosphere by spallation caused by cosmic-ray protons. Aluminium isotopes have found practical application in dating marine sediments, manganese nodules, glacial ice, quartz in rock exposures, and meteorites. The ratio of 26Al to 10Be has been used to study the role of transport, deposition, sediment storage, burial times, and erosion on 105 to 106 year time scales.[15] Cosmogenic 26Al was first applied in studies of the Moon and meteorites. Meteoroid fragments, after departure from their parent bodies, are exposed to intense cosmic-ray bombardment during their travel through space, causing substantial 26Al production. After falling to Earth, atmospheric shielding protects the meteorite fragments from further 26Al production, and its decay can then be used to determine the meteorite's terrestrial age. Meteorite research has also shown that 26Al was relatively abundant at the time of formation of our planetary system. Most meteoriticists believe that the energy released by the decay of 26Al was responsible for the melting and differentiation of some asteroids after their formation 4.55 billion years ago.[16]

Natural occurrence

In the Earth's crust, aluminium is the most abundant (8.3% by weight) metallic element and the third most abundant of all elements (after oxygen and silicon).[17] However, because of its strong affinity to oxygen, it is almost never found in the elemental state; instead it is found in oxides or silicates. Feldspars, the most common group of minerals in the earth's crust, are aluminosilicates. Native aluminium metal can be found as a minor phase in low oxygen fugacity environments, such as the interiors of certain volcanoes.[18] It also occurs in the minerals beryl, cryolite, garnet, spinel and turquoise.[17]Template:Inote Impurities in Al2O3, such as chromium or cobalt yield the gemstones ruby and sapphire, respectively.Template:Inote Pure Al2O3, known as Corundum, is one of the hardest materials known.[17]Template:Inote

Although aluminium is an extremely common and widespread element, the common aluminium minerals are not economic sources of the metal. Almost all metallic aluminium is produced from the ore bauxite (AlOx(OH)3-2x). Bauxite occurs as a weathering product of low iron and silica bedrock in tropical climatic conditions.[19] Large deposits of bauxite occur in Australia, Brazil, Guinea and Jamaica but the primary mining areas for the ore are in Ghana, Indonesia, Jamaica, Russia and Surinam.[20]Template:Inote Smelting of the ore mainly occurs in Australia, Brazil, Canada, Norway, Russia and the United States.Template:Inote

Production and refinement

Although aluminium is the most abundant metallic element in the Earth's crust (believed to be 7.5 to 8.1 percent), it is rare in its free form, occurring in oxygen-deficient environments such as volcanic mud, and it was once considered a precious metal more valuable than gold. Napoleon III, emperor of France, is reputed to have given a banquet where the most honoured guests were given aluminium utensils, while the other guests had to make do with gold.[21][22] The Washington Monument was completed, with the 100 ounce (2.8 kg) aluminium capstone being put in place on December 6, 1884, in an elaborate dedication ceremony. It was the largest single piece of aluminium cast at the time. At that time, aluminium was more expensive than silver, gold, or platinum. Aluminium has been produced in commercial quantities for just over 100 years.

Bauxite

Aluminium is a strongly reactive metal that forms a high-energy chemical bond with oxygen. Compared to most other metals, it is difficult to extract from ore, such as bauxite, due to the energy required to reduce aluminium oxide (Al2O3). For example, direct reduction with carbon, as is used to produce iron, is not chemically possible, since aluminium is a stronger reducing agent than carbon. Aluminium oxide has a melting point of about 2,000 °C. Therefore, it must be extracted by electrolysis. In this process, the aluminium oxide is dissolved in molten cryolite and then reduced to the pure metal. The operational temperature of the reduction cells is around 950 to 980 °C. Cryolite is found as a mineral in Greenland, but in industrial use it has been replaced by a synthetic substance. Cryolite is a chemical compound of aluminium, sodium, and calcium fluorides: (Na3AlF6). The aluminium oxide (a white powder) is obtained by refining bauxite in the Bayer process of Karl Bayer. (Previously, the Deville process was the predominant refining technology.)

The electrolytic process replaced the Wöhler process, which involved the reduction of anhydrous aluminium chloride with potassium. Both of the electrodes used in the electrolysis of aluminium oxide are carbon. Once the refined alumina is dissolved in the electrolyte, its ions are free to move around. The reaction at the cathode (negative electrode) is

Al3+ + 3 e → Al

Here the aluminium ion is being reduced (electrons are added). The aluminium metal then sinks to the bottom and is tapped off.

At the anode (positive electrode), oxygen is formed:

2 O2− → O2 + 4 e

This carbon anode is then oxidized by the oxygen, releasing carbon dioxide.

O2 + C → CO2

The anodes in a reduction cell must therefore be replaced regularly, since they are consumed in the process.

Unlike the anodes, the cathodes are not oxidized because there is no oxygen present, as the carbon cathodes are protected by the liquid aluminium inside the cells. Nevertheless, cathodes do erode, mainly due to electrochemical processes and metal movement. After five to ten years, depending on the current used in the electrolysis, a cell has to be rebuilt because of cathode wear.

World production trend of aluminium

Aluminium electrolysis with the Hall-Héroult process consumes a lot of energy, but alternative processes were always found to be less viable economically and/or ecologically. The worldwide average specific energy consumption is approximately 15±0.5 kilowatt-hours per kilogram of aluminium produced (52 to 56 MJ/kg). The most modern smelters achieve approximately 12.8 kW·h/kg (46.1 MJ/kg). (Compare this to the heat of reaction, 31 MJ/kg, and the Gibbs free energy of reaction, 29 MJ/kg.) Reduction line currents for older technologies are typically 100 to 200 kA; state-of-the-art smelters[23] operate at about 350 kA. Trials have been reported with 500 kA cells.

Recovery of the metal via recycling has become an important facet of the aluminium industry. Recycling involves melting the scrap, a process that requires only five percent of the energy used to produce aluminium from ore. However, a significant part (up to 15% of input material) is lost as dross (ash-like oxide).[24] Recycling was a low-profile activity until the late 1960s, when the growing use of aluminium beverage cans brought it to the public consciousness.

Electric power represents about 20% to 40% of the cost of producing aluminium, depending on the location of the smelter. Smelters tend to be situated where electric power is both plentiful and inexpensive, such as South Africa, Ghana, the South Island of New Zealand, Australia, the People's Republic of China, the Middle East, Russia, Quebec and British Columbia in Canada, and Iceland [citation needed] .

File:Aluminium output2.PNG
Aluminium output in 2005

In 2005, the People's Republic of China was the top producer of aluminium with almost a one-fifth world share, followed by Russia, Canada, and the USA, reports the British Geological Survey.

Over the last 50 years, Australia has become a major producer of bauxite ore and a major producer and exporter of alumina.[25] Australia produced 62 million tonnes of bauxite in 2005. The Australian deposits have some refining problems, some being high in silica but have the advantage of being shallow and relatively easy to mine.[26]

Chemistry

Oxidation state one

AlH is produced when aluminium is heated in an atmosphere of hydrogen. Al2O is made by heating the normal oxide, Al2O3, with silicon at 1800 °C in a vacuum.[citation needed]

Al2S can be made by heating Al2S3 with aluminium shavings at 1300 °C in a vacuum.[citation needed] It quickly disproportionates to the starting materials. The selenide is made in a parallel manner.

AlF, AlCl and AlBr exist in the gaseous phase when the tri-halide is heated with aluminium. Aluminium halides usually exist in the form AlX3. e.g. AlF3, AlCl3, AlBr3, AlI3 etc.

Oxidation state two

Aluminium monoxide, AlO, is present when aluminium powder burns in oxygen.

Oxidation state three

Fajans' rules show that the simple trivalent cation Al3+ is not expected to be found in anhydrous salts or binary compounds such as Al2O3. The hydroxide is a weak base and aluminium salts of weak acids, such as carbonate, can't be prepared. The salts of strong acids, such as nitrate, are stable and soluble in water, forming hydrates with at least six molecules of water of crystallization.

Aluminium hydride, (AlH3)n, can be produced from trimethylaluminium and an excess of hydrogen. It burns explosively in air. It can also be prepared by the action of aluminium chloride on lithium hydride in ether solution, but cannot be isolated free from the solvent. Alumino-hydrides of the most electropositive elements are known, the most useful being lithium aluminium hydride, Li[AlH4]. It decomposes into lithium hydride, aluminium and hydrogen when heated, and is hydrolysed by water. It has many uses in organic chemistry, particularly as a reducing agent. The aluminohalides have a similar structure.

Aluminium hydroxide may be prepared as a gelatinous precipitate by adding ammonia to an aqueous solution of an aluminium salt. It is amphoteric, being both a very weak acid, and forming aluminates with alkalis. It exists in various crystalline forms.

Aluminium carbide, Al4C3 is made by heating a mixture of the elements above 1000 °C. The pale yellow crystals have a complex lattice structure, and react with water or dilute acids to give methane. The acetylide, Al2(C2)3, is made by passing acetylene over heated aluminium.

Aluminium nitride, AlN, can be made from the elements at 800 °C. It is hydrolysed by water to form ammonia and aluminium hydroxide. Aluminium phosphide, AlP, is made similarly, and hydrolyses to give phosphine.

Aluminium oxide, Al2O3, occurs naturally as corundum, and can be made by burning aluminium in oxygen or by heating the hydroxide, nitrate or sulfate. As a gemstone, its hardness is only exceeded by diamond, boron nitride, and carborundum. It is almost insoluble in water. Aluminium sulfide, Al2S3, may be prepared by passing hydrogen sulfide over aluminium powder. It is polymorphic.

Aluminium iodide, AlI3, is a dimer with applications in organic synthesis. Aluminium fluoride, AlF3, is made by treating the hydroxide with HF, or can be made from the elements. It consists of a giant molecule which sublimes without melting at 1291 °C. It is very inert. The other trihalides are dimeric, having a bridge-like structure. Aluminium fluoride/water complexes: When aluminium and fluoride are together in aqueous solution, they readily form complex ions such as AlF(H2O)5+2, AlF3(H2O)30, AlF6-3. Of these, AlF6-3 is the most stable. This is explained by the fact that aluminium and fluoride, which are both very compact ions, fit together just right to form the octahedral aluminium hexafluoride complex. When aluminium and fluoride are together in water in a 1:6 molar ratio, AlF6-3 is the most common form, even in rather low concentrations.

Organo-metallic compounds of empirical formula AlR3 exist and, if not also giant molecules, are at least dimers or trimers. They have some uses in organic synthesis, for instance trimethylaluminium.

Applications

General use

A piece of aluminium metal.

Aluminium is the most widely used non-ferrous metal.[27] Global production of aluminium in 2005 was 31.9 million tonnes. It exceeded that of any other metal except iron (837.5 million tonnes).[28] Relatively pure aluminium is encountered only when corrosion resistance and/or workability is more important than strength or hardness. A thin layer of aluminium can be deposited onto a flat surface by physical vapor deposition or (very infrequently) chemical vapor deposition or other chemical means to form optical coatings and mirrors. When so deposited, a fresh, pure aluminium film serves as a good reflector (approximately 92%) of visible light and an excellent reflector (as much as 98%) of medium and far infrared.

Pure aluminium has a low tensile strength, but when combined with thermo-mechanical processing, aluminium alloys display a marked improvement in mechanical properties, especially when tempered. Aluminium alloys form vital components of aircraft and rockets as a result of their high strength-to-weight ratio. Aluminium readily forms alloys with many elements such as copper, zinc, magnesium, manganese and silicon (e.g., duralumin). Today, almost all bulk metal materials that are referred to loosely as "aluminium," are actually alloys. For example, the common aluminium foils are alloys of 92% to 99% aluminium.[29]

Some of the many uses for aluminium metal are in:

household aluminium foil

Aluminium compounds

  • Aluminium hydroxide (Al(OH)3) is used: as an antacid, as a mordant, in water purification, in the manufacture of glass and ceramic and in the waterproofing of fabrics.
  • Aluminium sulfate (Al2(SO4)3) is used: in the manufacture of paper, as a mordant, in a fire extinguisher, in water purification and sewage treatment, as a food additive, in fireproofing, and in leather tanning.
  • In many vaccines, certain aluminium salts serve as an immune adjuvant (immune response booster) to allow the protein in the vaccine to achieve sufficient potency as an immune stimulant.

Aluminium alloys in structural applications

Aluminium foam

Aluminium alloys with a wide range of properties are used in engineering structures. Alloy systems are classified by a number system (ANSI) or by names indicating their main alloying constituents (DIN and ISO).

The strength and durability of aluminium alloys vary widely, not only as a result of the components of the specific alloy, but also as a result of heat treatments and manufacturing processes. A lack of knowledge of these aspects has from time to time led to improperly designed structures and gained aluminium a bad reputation. (See main article)

One important structural limitation of aluminium alloys is their fatigue strength. Unlike steels, aluminium alloys have no well-defined fatigue limit, meaning that fatigue failure will eventually occur under even very small cyclic loadings. This implies that engineers must assess these loads and design for a fixed life rather than an infinite life.

Another important property of aluminium alloys is their sensitivity to heat. Workshop procedures involving heating are complicated by the fact that aluminium, unlike steel, will melt without first glowing red. Forming operations where a blow torch is used therefore requires some expertise, since no visual signs reveal how close the material is to melting. Aluminium alloys, like all structural alloys, also are subject to internal stresses following heating operations such as welding and casting. The problem with aluminium alloys in this regard is their low melting point, which make them more susceptible to distortions from thermally induced stress relief. Controlled stress relief can be done during manufacturing by heat-treating the parts in an oven, followed by gradual cooling -- in effect annealing the stresses.

The low melting point of aluminium alloys has not precluded their use in rocketry; even for use in constructing combustion chambers where gases can reach 3500 K. The Agena upper stage engine used a regeneratively cooled aluminium design for some parts of the nozzle, including the thermally critical throat region.

Household wiring

Compared to copper, aluminium has about 65% of the electrical conductivity by volume, although 200% by weight. Traditionally copper is used as household wiring material. In the 1960s aluminium was considerably cheaper than copper, and so was introduced for household electrical wiring in the United States, even though many fixtures had not been designed to accept aluminium wire. However, in some cases the greater coefficient of thermal expansion of aluminium causes the wire to expand and contract relative to the dissimilar metal screw connection, eventually loosening the connection. Also, pure aluminium has a tendency to creep under steady sustained pressure (to a greater degree as the temperature rises), again loosening the connection. Finally, Galvanic corrosion from the dissimilar metals increased the electrical resistance of the connection.

All of this resulted in overheated and loose connections, and this in turn resulted in fires. Builders then became wary of using the wire, and many jurisdictions outlawed its use in very small sizes in new construction. Eventually, newer fixtures were introduced with connections designed to avoid loosening and overheating. At first they were marked "Al/Cu", but they now bear a "CO/ALR" coding. In older assemblies, workers forestall the heating problem using a properly-done crimp of the aluminium wire to a short "pigtail" of copper wire. Today, new alloys, designs, and methods are used for aluminium wiring in combination with aluminium terminations.

History

The statue of the Anteros as the Angel of Christian Charity (commonly mistaken for Eros) in Piccadilly Circus London, was made in 1893 and is one of the first statues to be cast in aluminium.

Ancient Greeks and Romans used aluminium salts as dyeing mordants and as astringents for dressing wounds; alum is still used as a styptic. In 1761 Guyton de Morveau suggested calling the base alum alumine. In 1808, Humphry Davy identified the existence of a metal base of alum, which he at first termed alumium and later aluminum (see Etymology section, below).

The metal was first produced in 1825 (in an impure form) by Danish physicist and chemist Hans Christian Ørsted. He reacted anhydrous aluminium chloride with potassium amalgam and yielded a lump of metal looking similar to tin.[31] Friedrich Wöhler was aware of these experiments and cited them, but after redoing the experiments of Ørsted he concluded that this metal was pure potassium. He conducted a similar experiment in 1827 by mixing anhydrous aluminium chloride with potassium and yielded aluminium.[31] Wöhler is generally credited with isolating aluminium (Latin alumen, alum), but also Ørsted can be listed as its discoverer.[32] Further, Pierre Berthier discovered aluminium in bauxite ore and successfully extracted it.[33] Frenchman Henri Etienne Sainte-Claire Deville improved Wöhler's method in 1846, and described his improvements in a book in 1859, chief among these being the substitution of sodium for the considerably more expensive potassium.

(Note: The title of Deville's book is De l'aluminium, ses propriétés, sa fabrication (Paris, 1859). Deville likely also conceived the idea of the electrolysis of aluminium oxide dissolved in cryolite; however, Charles Martin Hall and Paul Héroult might have developed the more practical process after Deville.)

Before the Hall-Héroult process was developed, aluminium was exceedingly difficult to extract from its various ores. This made pure aluminium more valuable than gold. Bars of aluminium were exhibited alongside the French crown jewels at the Exposition Universelle of 1855, and Napoleon III was said to have reserved a set of aluminium dinner plates for his most honoured guests.

Aluminium was selected as the material to be used for the apex of the Washington Monument in 1884, a time when one ounce (30 grams) cost the daily wage of a common worker on the project;[34] aluminium was about the same value as silver.

The Cowles companies supplied aluminium alloy in quantity in the United States and England using smelters like the furnace of Carl Wilhelm Siemens by 1886.[35] Charles Martin Hall of Ohio in the U.S. and Paul Héroult of France independently developed the Hall-Héroult electrolytic process that made extracting aluminium from minerals cheaper and is now the principal method used worldwide. The Hall-Heroult process cannot produce Super Purity Aluminium directly. Hall's process,[36] in 1888 with the financial backing of Alfred E. Hunt, started the Pittsburgh Reduction Company today known as Alcoa. Héroult's process was in production by 1889 in Switzerland at Aluminium Industrie, now Alcan, and at British Aluminium, now Luxfer Group and Alcoa, by 1896 in Scotland.[37]

By 1895 the metal was being used as a building material as far away as Sydney, Australia in the dome of the Chief Secretary's Building.

Many navies use an aluminium superstructure for their vessels, however, the 1975 fire aboard USS Belknap that gutted her aluminium superstructure, as well as observation of battle damage to British ships during the Falklands War, led to many navies switching to all steel superstructures. The Arleigh Burke class was the first such U.S. ship, being constructed entirely of steel.

In 2008 the price of aluminium peaked at $1.45/lb in July but dropped to $0.7/lb by December.[38]

Etymology

Nomenclature history

The earliest citation given in the Oxford English Dictionary for any word used as a name for this element is alumium, which British chemist and inventor Humphry Davy employed in 1808 for the metal he was trying to isolate electrolytically from the mineral alumina. The citation is from his journal Philosophical Transactions: "Had I been so fortunate as..to have procured the metallic substances I was in search of, I should have proposed for them the names of silicium, alumium, zirconium, and glucium."[39]

By 1812, Davy had settled on aluminum. He wrote in the journal Chemical Philosophy: "As yet Aluminum has not been obtained in a perfectly free state."[40] But the same year, an anonymous contributor to the Quarterly Review, a British political-literary journal, objected to aluminum and proposed the name aluminium, "for so we shall take the liberty of writing the word, in preference to aluminum, which has a less classical sound."[41]

The -ium suffix had the advantage of conforming to the precedent set in other newly discovered elements of the time: potassium, sodium, magnesium, calcium, and strontium (all of which Davy had isolated himself). Nevertheless, -um spellings for elements were not unknown at the time, as for example platinum, known to Europeans since the sixteenth century, molybdenum, discovered in 1778, and tantalum, discovered in 1802.

Americans adopted -ium to fit the standard form of the periodic table of elements, for most of the nineteenth century, with aluminium appearing in Webster's Dictionary of 1828. In 1892, however, Charles Martin Hall used the -um spelling in an advertising handbill for his new electrolytic method of producing the metal, despite his constant use of the -ium spelling in all the patents[36] he filed between 1886 and 1903.[42] It has consequently been suggested that the spelling reflects an easier to pronounce word with one fewer syllable, or that the spelling on the flier was a mistake. Hall's domination of production of the metal ensured that the spelling aluminum became the standard in North America; the Webster Unabridged Dictionary of 1913, though, continued to use the -ium version.

In 1926, the American Chemical Society officially decided to use aluminum in its publications; American dictionaries typically label the spelling aluminium as a British variant.

Present-day spelling

In the UK and most other countries using British spelling, only aluminium (with an i before -um) is used. In the United States, the spelling aluminium is largely unknown, and the spelling aluminum predominates.[43][44] The Canadian Oxford Dictionary prefers aluminum, whereas the Australian Macquarie Dictionary prefers aluminium.

The International Union of Pure and Applied Chemistry (IUPAC) adopted aluminium as the standard international name for the element in 1990, but three years later recognized aluminum as an acceptable variant. Hence their periodic table includes both, but places aluminium first in alphabetical order.[45] IUPAC officially prefers the use of aluminium in its internal publications, although several IUPAC publications use the spelling aluminum.[46]

Health concerns

Despite its natural abundance, aluminium has no known function in living cells and presents some toxic effects in elevated concentrations. Its toxicity can be traced to deposition in bone and the central nervous system, which is particularly increased in patients with reduced renal function. Because aluminium competes with calcium for absorption, increased amounts of dietary aluminium may contribute to the reduced skeletal mineralization (osteopenia) observed in preterm infants and infants with growth retardation. In very high doses, aluminium can cause neurotoxicity, and is associated with altered function of the blood-brain barrier.[47] A small percentage of people are allergic to aluminium and experience contact dermatitis, digestive disorders, vomiting or other symptoms upon contact or ingestion of products containing aluminum, such as deodorants or antacids. In other people aluminium is not considered as toxic as heavy metals, but there is evidence of some toxicity if it is consumed in excessive amounts[citation needed]. Although the use of aluminium cookware has not been shown to lead to aluminium toxicity in general, excessive consumption of antacids containing aluminium compounds and excessive use of aluminium-containing antiperspirants provide more significant exposure levels. Studies have shown that consumption of acidic foods or liquids with aluminium significantly increases aluminium absorption,[48] and maltol has been shown to increase the accumulation of aluminium in nervous and osseus tissue.[49] Furthermore, aluminium increases estrogen-related gene expression in human breast cancer cells cultured in the laboratory.[50] These salts' estrogen-like effects have led to their classification as a metalloestrogen.

Because of its potentially toxic effects, aluminium's use in some antiperspirants and food additives is controversial. Although there is little evidence that normal exposure to aluminium represents a risk to healthy adults[51], several studies point to risks associated with increased exposure to the metal. Aluminium in food may be absorbed more than aluminium from water.[52] Some researchers have expressed concerns that the aluminium in antiperspirants may increase the risk of breast cancer,[53] and aluminium has controversially been implicated as a factor in Alzheimer's disease.[54]

According to The Alzheimer's Society, the overwhelming medical and scientific opinion is that studies have not convincingly demonstrated a causal relationship between aluminium and Alzheimer's disease.[55] Nevertheless, some studies cite aluminium exposure as a risk factor for Alzheimer's disease, as some brain plaques have been found to contain increased levels of the metal. Research in this area has been inconclusive; aluminium accumulation may be a consequence of the disease rather than a causal agent. In any event, if there is any toxicity of aluminium, it must be via a very specific mechanism, since total human exposure to the element in the form of naturally occurring clay in soil and dust is enormously large over a lifetime.[56][57] Scientific consensus does not yet exist about whether aluminium exposure could directly increase the risk of Alzheimer's disease.[55]

Effect on plants

Aluminium is primary among the factors that reduce plant growth on acid soils.[58][59][60] Although it is generally harmless to plant growth in pH-neutral soils, the concentration in acid soils of toxic Al3+ cations increases and disturbs root growth and function.[citation needed]

Most acid soils are saturated with aluminium rather than hydrogen ions. The acidity of the soil is therefore a result of hydrolysis of aluminium.[61] This concept of "corrected lime potential"[62] to define the degree of base saturation in soils became the basis for procedures now used in soil testing laboratories to determine the "lime requirement" of soils.[63]

Wheat's adaptation to allow aluminium tolerance is such that the aluminium induces a release of organic compounds that bind to the harmful aluminium cations.[citation needed] Sorghum is believed to have the same tolerance mechanism. The first gene for aluminium tolerance has been identified in wheat. A group in the U.S. Department of Agriculture showed that sorghum's aluminium tolerance is controlled by a single gene, as for wheat.[citation needed] This is not the case in all plants.

See also

References

  1. ^ "aluminum". Oxford English Dictionary (Online ed.). Oxford University Press. (Subscription or participating institution membership required.)
  2. ^ "Standard Atomic Weights: Aluminium". CIAAW. 2017.
  3. ^ Prohaska, Thomas; Irrgeher, Johanna; Benefield, Jacqueline; Böhlke, John K.; Chesson, Lesley A.; Coplen, Tyler B.; Ding, Tiping; Dunn, Philip J. H.; Gröning, Manfred; Holden, Norman E.; Meijer, Harro A. J. (2022-05-04). "Standard atomic weights of the elements 2021 (IUPAC Technical Report)". Pure and Applied Chemistry. doi:10.1515/pac-2019-0603. ISSN 1365-3075.
  4. ^ a b c Arblaster, John W. (2018). Selected Values of the Crystallographic Properties of Elements. Materials Park, Ohio: ASM International. ISBN 978-1-62708-155-9.
  5. ^ Unstable carbonyl of Al(0) has been detected in reaction of Al2(CH3)6 with carbon monoxide; see Sanchez, Ramiro; Arrington, Caleb; Arrington Jr., C. A. (December 1, 1989). "Reaction of trimethylaluminum with carbon monoxide in low-temperature matrixes". American Chemical Society. 111 (25): 9110-9111. doi:10.1021/ja00207a023. OSTI 6973516.
  6. ^ Dohmeier, C.; Loos, D.; Schnöckel, H. (1996). "Aluminum(I) and Gallium(I) Compounds: Syntheses, Structures, and Reactions". Angewandte Chemie International Edition. 35 (2): 129–149. doi:10.1002/anie.199601291.
  7. ^ Tyte, D. C. (1964). "Red (B2Π–A2σ) Band System of Aluminium Monoxide". Nature. 202 (4930): 383. Bibcode:1964Natur.202..383T. doi:10.1038/202383a0. S2CID 4163250.
  8. ^ Lide, D. R. (2000). "Magnetic susceptibility of the elements and inorganic compounds" (PDF). CRC Handbook of Chemistry and Physics (81st ed.). CRC Press. ISBN 0849304814.
  9. ^ Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S.; Audi, G. (2021). "The NUBASE2020 evaluation of nuclear properties" (PDF). Chinese Physics C. 45 (3): 030001. doi:10.1088/1674-1137/abddae.
  10. ^ Mougeot, X. (2019). "Towards high-precision calculation of electron capture decays". Applied Radiation and Isotopes. 154 (108884). doi:10.1016/j.apradiso.2019.108884.
  11. ^ Bassam Z. Shakhashiri. "Chemical of the Week: Aluminum". Science is Fun. Retrieved 2007-08-28.
  12. ^ a b Polmear, I. J. (1995). Light Alloys: Metallurgy of the Light Metals. Arnold. ISBN 9780340632079.
  13. ^ G. E. Dieter, Mechanical Metallurgy, McGraw-Hill, 1988
  14. ^ John F. Cochran and D. E. Mapother (1958). "Superconducting Transition in Aluminum". Physical Review. 111 (1): 132–142. doi:10.1103/PhysRev.111.132. {{cite journal}}: Unknown parameter |month= ignored (help)
  15. ^ Cosmogenic Isotopes and Aluminum
  16. ^ Robert T. Dodd, Thunderstones and Shooting Stars, pp. 89-90. ISBN 0-674-89137-6.
  17. ^ a b c Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. ISBN 978-0-08-037941-8.Template:Inote
  18. ^ Aluminum Mineral Data, retrieved 2008-07-09
  19. ^ Guilbert, John M. and Carles F. Park, The Geology of Ore Deposits, Freeman, 1986, pp. 774-795 ISBN 0-7167-1456-6
  20. ^ Emsley, John (2001). "Aluminium". Nature's Building Blocks: An A-Z Guide to the Elements. Oxford, England, UK: Oxford University Press. ISBN 0198503407.{{cite book}}: CS1 maint: ref duplicates default (link)
  21. ^ S Venetski (1969). ""Silver" from clay". Metallurgist. 13 (7): 451–453. doi:10.1007/BF00741130. {{cite journal}}: Unknown parameter |month= ignored (help)
  22. ^ ChemMatters October 1990 Page 14
  23. ^ "Aluminium Smelters". AME Mineral Economics. Retrieved 2008-04-17.
  24. ^ "Benefits of Recycling". Ohio Department of Natural Resources.
  25. ^ "The Australian Industry". Australian Aluminium Council. Retrieved 2007-08-11.
  26. ^ "Australian Bauxite". Australian Aluminium Council. Retrieved 2007-08-11.
  27. ^ "aluminum". Encyclopædia Britannica.
  28. ^ L E Hetherington, T J Brown, A J Benham, P A J Lusty, N E Idoine (2007). World Mineral Production: 2001 - 2005 (available online). British Geological Survey. ISBN 978-0-85272-592-4.{{cite book}}: CS1 maint: multiple names: authors list (link)
  29. ^ L. S. Millberg. "Aluminum Foil". How Products are Made. Retrieved 2007-08-11.
  30. ^ Aluminum in Watchmaking
  31. ^ a b Wöhler, Friedrich. "Ueber das Aluminium". Annalen der Physik und Chemie.
  32. ^ Yinon Bentor. "Periodic Table: Aluminum". ChemicalElements.com. Retrieved 2007-08-11.
  33. ^ "Pierre Berthier". Today in Science History. Retrieved 2007-08-11.
  34. ^ George J. Binczewski (1995). "The Point of a Monument: A History of the Aluminum Cap of the Washington Monument". JOM. 47 (11): 20–25.
  35. ^ "Cowles' Aluminium Alloys". The Manufacturer and Builder. 18 (1). New York: Western and Company, via Cornell University Library: 13. 1886. Retrieved 2007-10-27. {{cite journal}}: Unknown parameter |month= ignored (help) and McMillan, Walter George (1891). A Treatise on Electro-Metallurgy. London, Philadelphia: Charles Griffin and Company, J.B. Lippincott Company, via Google Books scan of New York Public Library copy. pp. 302–305. Retrieved 2007-10-26. and Sackett, William Edgar, John James Scannell and Mary Eleanor Watson (1917/1918). New Jersey's First Citizens. New Jersey: J.J. Scannell via Google Books scan of New York Public Library copy. pp. 103–105. Retrieved 2007-10-25. {{cite book}}: Check date values in: |date= (help)CS1 maint: multiple names: authors list (link)
  36. ^ a b US patent 400664, Charles Martin Hall, "Process of Reducing Aluminium from its Fluoride Salts by Electrolysis", issued 1889-04-02 
  37. ^ Donald Holmes Wallace (1977) [1937]. Market Control in the Aluminum Industry. Harvard University Press via Ayer Publishing via Google Books limited view. p. 6. ISBN 0-4050-9786-7. Retrieved 2007-10-27.
  38. ^ Aluminum prices
  39. ^ "alumium", Oxford English Dictionary. Ed. J.A. Simpson and E.S.C. Weiner, second edition Oxford: Clarendon Press, 1989. OED Online Oxford University Press. Accessed 29 October 2006. Citation is listed as "1808 SIR H. DAVY in Phil. Trans. XCVIII. 353". The ellipsis in the quotation is as it appears in the OED citation.
  40. ^ "aluminum", Oxford English Dictionary. Ed. J.A. Simpson and E.S.C. Weiner, second edition Oxford: Clarendon Press, 1989. OED Online Oxford University Press. Accessed 29 October 2006. Citation is listed as "1812 SIR H. DAVY Chem. Philos. I. 355"
  41. ^ "aluminium", Oxford English Dictionary. Ed. J.A. Simpson and E.S.C. Weiner, second edition Oxford: Clarendon Press, 1989. OED Online Oxford University Press. Accessed 29 October 2006. Citation is listed as "1812 Q. Rev. VIII. 72"
  42. ^ Peter Meiers. "Manufacture of Aluminum". The History of Fluorine, Fluoride and Fluoridation.
  43. ^ Greenwood, Norman N.; Earnshaw, Alan (1984). Chemistry of the Elements. Oxford: Pergamon Press. p. 243. ISBN 978-0-08-022057-4.
  44. ^ John Bremner, Words on Words: A Dictionary for Writers and Others Who Care about Words, page 22–23. ISBN 0-231-04493-3
  45. ^ IUPAC Periodic Table of the Elements
  46. ^ IUPAC Web site publication search for 'aluminum'
  47. ^ Banks, W.A. (1989). "Aluminum-induced neurotoxicity: alterations in membrane function at the blood-brain barrier". Neurosci Biobehav Rev. 13 (1): 47–53. doi:10.1016/S0149-7634(89)80051-X. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  48. ^ Slanina, P. (1986). "Dietary citric acid enhances absorption of aluminum in antacids". Clinical Chemistry. 32. American Association for Clinical Chemistry: 539–541. Retrieved 2008-10-09. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  49. ^ van Ginkel MF, van der Voet GB, D'Haese PC, De Broe ME, de Wolff FA, Effect of citric acid and maltol on the accumulation of aluminum in rat brain and bone, J Lab Clin Med. 1993 Mar;121: 453-60, PMID 8445293
  50. ^ Metalloestrogens: an emerging class of inorganic xenoestrogens with potential to add to the oestrogenic burden of the human breast J Appl Toxicol. 2006 May-Jun;26(3):191-7
  51. ^ H. J. Gitelman, "Physiology of Aluminum in Man," in Aluminum and Health, CRC Press, 1988, ISBN 0824780264
  52. ^ Yokel RA, Hicks CL, Florence RL (2008). "Aluminum bioavailability from basic sodium aluminum phosphate, an approved food additive emulsifying agent, incorporated in cheese". Food and chemical toxicology. 46 (6): 2261–6. doi:10.1016/j.fct.2008.03.004. PMID 18436363. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  53. ^ Exley C, Charles LM, Barr L, Martin C, Polwart A, Darbre PD (2007). "Aluminium in human breast tissue". J. Inorg. Biochem. 101 (9): 1344–6. doi:10.1016/j.jinorgbio.2007.06.005. PMID 17629949. {{cite journal}}: Unknown parameter |month= ignored (help)CS1 maint: multiple names: authors list (link)
  54. ^ Ferreira PC, Piai Kde A, Takayanagui AM, Segura-Muñoz SI (2008). "Aluminum as a risk factor for Alzheimer's disease". Rev Lat Am Enfermagem. 16 (1): 151–7. PMID 18392545.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  55. ^ a b Aluminium and Alzheimer's disease, The Alzheimer's Society, accessed 30 January 2009.
  56. ^ "Alzheimer's Disease and Aluminum". National Institute of Environmental Health Sciences. 2005. {{cite web}}: Unknown parameter |month= ignored (help)
  57. ^ Michael Hopkin (21 April 2006). "Death of Alzheimer victim linked to aluminium pollution". news @ nature.com. doi:10.1038/news060417-10.
  58. ^ Andersson, Maud (1988). "Toxicity and tolerance of aluminium in vascular plants". Water, Air, & Soil Pollution. 39 (3–4): 439–462. doi:10.1007/BF00279487. {{cite journal}}: Cite has empty unknown parameter: |coauthors= (help)
  59. ^ Horst, Walter J. (1995). "The role of the apoplast in aluminium toxicity and resistance of higher plants: A review". Zeitschrift für Pflanzenernährung und Bodenkunde. 158 (5): 419–428. doi:10.1002/jpln.19951580503. {{cite journal}}: Cite has empty unknown parameter: |coauthors= (help)
  60. ^ Ma, Jian Feng (2001). "Aluminium tolerance in plants and the complexing role of organic acids". Trends in Plant Science. 6 (6): 273–278. doi:10.1016/S1360-1385(01)01961-6. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  61. ^ Turner, R.C. and Clark J.S., 1966, Lime potential in acid clay and soil suspensions. Trans. Comm. II & IV Int. Soc. Soil Science, pp. 208-215
  62. ^ http://scholar.google.ca/scholar?q=info:VT442wAXDoAJ:scholar.google.com/&hl=en&lr=&output=viewport&shm=1 "A Study of the Lime Potential, R.C. Turner, Research Branch, Department Of Agriculture, 1965"
  63. ^ Applying lime to soils reduces the Aluminum toxicity to plants. "One Hundred Harvests Research Branch Agriculture Canada 1886-1986". Historical series / Agriculture Canada - Série historique / Agriculture Canada. Government of Canada. Retrieved 2008-12-22. {{cite web}}: Cite has empty unknown parameter: |coauthors= (help) Note this link loads slowly

Template:Link FA
Cite error: There are <ref group=lower-alpha> tags or {{efn}} templates on this page, but the references will not show without a {{reflist|group=lower-alpha}} template or {{notelist}} template (see the help page).