Tin

From Wikipedia, the free encyclopedia
Jump to: navigation, search
This article is about the chemical element. For other uses, see Tin (disambiguation).
Tin,  50Sn
Sn-Alpha-Beta.jpg
left: white, beta, β; right: gray, alpha, α
General properties
Name, symbol tin, Sn
Appearance silvery-white (beta, β) or gray (alpha, α)
Allotropes alpha, α (gray); beta, β (white)
Pronunciation /ˈtɪn/
TIN
Tin in the periodic table
Hydrogen (diatomic nonmetal)
Helium (noble gas)
Lithium (alkali metal)
Beryllium (alkaline earth metal)
Boron (metalloid)
Carbon (polyatomic nonmetal)
Nitrogen (diatomic nonmetal)
Oxygen (diatomic nonmetal)
Fluorine (diatomic nonmetal)
Neon (noble gas)
Sodium (alkali metal)
Magnesium (alkaline earth metal)
Aluminium (post-transition metal)
Silicon (metalloid)
Phosphorus (polyatomic nonmetal)
Sulfur (polyatomic nonmetal)
Chlorine (diatomic nonmetal)
Argon (noble gas)
Potassium (alkali metal)
Calcium (alkaline earth metal)
Scandium (transition metal)
Titanium (transition metal)
Vanadium (transition metal)
Chromium (transition metal)
Manganese (transition metal)
Iron (transition metal)
Cobalt (transition metal)
Nickel (transition metal)
Copper (transition metal)
Zinc (transition metal)
Gallium (post-transition metal)
Germanium (metalloid)
Arsenic (metalloid)
Selenium (polyatomic nonmetal)
Bromine (diatomic nonmetal)
Krypton (noble gas)
Rubidium (alkali metal)
Strontium (alkaline earth metal)
Yttrium (transition metal)
Zirconium (transition metal)
Niobium (transition metal)
Molybdenum (transition metal)
Technetium (transition metal)
Ruthenium (transition metal)
Rhodium (transition metal)
Palladium (transition metal)
Silver (transition metal)
Cadmium (transition metal)
Indium (post-transition metal)
Tin (post-transition metal)
Antimony (metalloid)
Tellurium (metalloid)
Iodine (diatomic nonmetal)
Xenon (noble gas)
Caesium (alkali metal)
Barium (alkaline earth metal)
Lanthanum (lanthanide)
Cerium (lanthanide)
Praseodymium (lanthanide)
Neodymium (lanthanide)
Promethium (lanthanide)
Samarium (lanthanide)
Europium (lanthanide)
Gadolinium (lanthanide)
Terbium (lanthanide)
Dysprosium (lanthanide)
Holmium (lanthanide)
Erbium (lanthanide)
Thulium (lanthanide)
Ytterbium (lanthanide)
Lutetium (lanthanide)
Hafnium (transition metal)
Tantalum (transition metal)
Tungsten (transition metal)
Rhenium (transition metal)
Osmium (transition metal)
Iridium (transition metal)
Platinum (transition metal)
Gold (transition metal)
Mercury (transition metal)
Thallium (post-transition metal)
Lead (post-transition metal)
Bismuth (post-transition metal)
Polonium (post-transition metal)
Astatine (metalloid)
Radon (noble gas)
Francium (alkali metal)
Radium (alkaline earth metal)
Actinium (actinide)
Thorium (actinide)
Protactinium (actinide)
Uranium (actinide)
Neptunium (actinide)
Plutonium (actinide)
Americium (actinide)
Curium (actinide)
Berkelium (actinide)
Californium (actinide)
Einsteinium (actinide)
Fermium (actinide)
Mendelevium (actinide)
Nobelium (actinide)
Lawrencium (actinide)
Rutherfordium (transition metal)
Dubnium (transition metal)
Seaborgium (transition metal)
Bohrium (transition metal)
Hassium (transition metal)
Meitnerium (unknown chemical properties)
Darmstadtium (unknown chemical properties)
Roentgenium (unknown chemical properties)
Copernicium (transition metal)
Ununtrium (unknown chemical properties)
Flerovium (post-transition metal)
Ununpentium (unknown chemical properties)
Livermorium (unknown chemical properties)
Ununseptium (unknown chemical properties)
Ununoctium (unknown chemical properties)
Ge

Sn

Pb
indiumtinantimony
Atomic number (Z) 50
Element category   post-transition metal
Group, block group 14 (carbon group), p-block
Period period 5
Standard atomic weight (±) (Ar) 118.710(7)[1]
Electron configuration [Kr] 4d10 5s2 5p2
per shell
2, 8, 18, 18, 4
Physical properties
Phase solid
Melting point 505.08 K ​(231.93 °C, ​449.47 °F)
Boiling point 2875 K ​(2602 °C, ​4716 °F)
Density near r.t. white, β: 7.365 g/cm3
gray, α: 5.769 g/cm3
when liquid, at m.p. 6.99 g/cm3
Heat of fusion white, β: 7.03 kJ/mol
Heat of vaporization white, β: 296.1 kJ/mol
Molar heat capacity white, β: 27.112 J/(mol·K)
vapor pressure
P (Pa) 1 10 100 1 k 10 k 100 k
at T (K) 1497 1657 1855 2107 2438 2893
Atomic properties
Oxidation states 4, 3,[2] 2, 1,[3] −1, −2, −3, −4 ​(an amphoteric oxide)
Electronegativity Pauling scale: 1.96
Ionization energies 1st: 708.6 kJ/mol
2nd: 1411.8 kJ/mol
3rd: 2943.0 kJ/mol
Atomic radius empirical: 140 pm
Covalent radius 139±4 pm
Van der Waals radius 217 pm
Miscellanea
Crystal structure tetragonal
Tetragonal crystal structure for tin

white (β)
Crystal structure diamond cubic
Diamond cubic crystal structure for tin

gray (α)
Speed of sound thin rod 2730 m/s (at r.t.) (rolled)
Thermal expansion 22.0 µm/(m·K) (at 25 °C)
Thermal conductivity 66.8 W/(m·K)
Electrical resistivity 115 nΩ·m (at 0 °C)
Magnetic ordering gray: diamagnetic[4]
white (β): paramagnetic
Young's modulus 50 GPa
Shear modulus 18 GPa
Bulk modulus 58 GPa
Poisson ratio 0.36
Brinell hardness 50–440 MPa
CAS Number 7440-31-5
History
Discovery around 3500 BC
Most stable isotopes of tin
iso NA half-life DM DE (MeV) DP
112Sn 0.97% (β+β+) 1.9222 112Cd
114Sn 0.66% (SF) <27.965
115Sn 0.34% (SF) <26.791
116Sn 14.54% (SF) <25.905
117Sn 7.68% (SF) <25.334
118Sn 24.22% (SF) <23.815
119Sn 8.59% (SF) <23.140
120Sn 32.58% (SF) <21.824
122Sn 4.63% (ββ) 0.3661 122Te
124Sn 5.79% >1×1017 y β) 2.2870 124Te
126Sn trace 2.3×105 y β 0.380 + 126Sb
Decay modes in parentheses are predicted, but have not yet been observed
· references

Tin is a chemical element with the symbol Sn (for Latin: stannum) and atomic number 50. It is a main group metal in group 14 of the periodic table. Tin shows a chemical similarity to both neighboring group-14 elements, germanium and lead, and has two possible oxidation states, +2 and the slightly more stable +4. Tin is the 49th most abundant element and has, with 10 stable isotopes, the largest number of stable isotopes in the periodic table. It is a silvery, malleable other metal that is not easily oxidized in air, obtained chiefly from the mineral cassiterite where it occurs as tin dioxide, SnO2.

The first alloy used on a large scale since 3000 BC was bronze, an alloy of tin and copper. After 600 BC, pure metallic tin was produced. Pewter, which is an alloy of 85–90% tin with the remainder commonly consisting of copper, antimony and lead, was used for flatware from the Bronze Age until the 20th century. In modern times, tin is used in many alloys, most notably tin/lead soft solders, which are typically 60% or more tin. Another large application for tin is corrosion-resistant tin plating of steel. Because of its low toxicity, tin-plated metal was used for food packaging as tin cans, which are now made mostly of steel,[citation needed] even though the name is kept in English.

Characteristics[edit]

Physical properties[edit]

Droplet of solidified molten tin

Tin is a malleable, ductile and highly crystalline silvery-white metal. When a bar of tin is bent, a crackling sound known as the tin cry can be heard due to the twinning of the crystals.[5] Tin melts at a low temperature of about 232 °C (450 °F), which is further reduced to 177.3 °C (351.1 °F) for 11 nm particles.[6]

β-tin (the metallic form, or white tin), which is stable at and above room temperature, is malleable. In contrast, α-tin (nonmetallic form, or gray tin), which is stable below 13.2 °C (55.8 °F), is brittle. α-tin has a diamond cubic crystal structure, similar to diamond, silicon or germanium. α-tin has no metallic properties at all because its atoms form a covalent structure where electrons cannot move freely. It is a dull-gray powdery material with no common uses, other than a few specialized semiconductor applications.[5] These two allotropes, α-tin and β-tin, are more commonly known as gray tin and white tin, respectively. Two more allotropes, γ and σ, exist at temperatures above 161 °C (322 °F)  and pressures above several GPa.[7] In cold conditions, β-tin tends to transform spontaneously into α-tin, a phenomenon known as "tin pest".[8] Although the α-β transformation temperature is nominally 13.2 °C (55.8 °F), impurities (e.g. Al, Zn, etc.) lower the transition temperature well below 0 °C (32 °F) and, on the addition of Sb or Bi, the transformation may not occur at all, increasing the durability of the tin.[9]

Commercial grades of tin (99.8%) resist transformation because of the inhibiting effect of the small amounts of bismuth, antimony, lead and silver present as impurities. Alloying elements such as copper, antimony, bismuth, cadmium and silver increase its hardness. Tin tends rather easily to form hard, brittle intermetallic phases, which are often undesirable. It does not form wide solid solution ranges in other metals in general, and there are few elements that have appreciable solid solubility in tin. Simple eutectic systems, however, occur with bismuth, gallium, lead, thallium and zinc.[9]

Tin becomes a superconductor below 3.72 K.[10] In fact, tin was one of the first superconductors to be studied; the Meissner effect, one of the characteristic features of superconductors, was first discovered in superconducting tin crystals.[11]

Chemical properties[edit]

Tin resists corrosion from water, but can be attacked by acids and alkalis. Tin can be highly polished and is used as a protective coat for other metals.[5] In this case a protective oxide layer prevents further oxidation. This oxide layer forms on pewter and other tin alloys.[12] Tin acts as a catalyst when oxygen is in solution and helps accelerate chemical attack.[clarification needed][citation needed][5]

Isotopes[edit]

Main article: Isotopes of tin

Tin has ten stable isotopes, with atomic masses of 112, 114 through 120, 122 and 124, the greatest number of any element. Of these, the most abundant ones are 120Sn (at almost a third of all tin), 118Sn, and 116Sn, while the least abundant one is 115Sn. The isotopes possessing even mass numbers have no nuclear spin, while the odd ones have a spin of +1/2. Tin, with its three common isotopes 116Sn, 118Sn and 120Sn, is among the easiest elements to detect and analyze by NMR spectroscopy, and its chemical shifts are referenced against SnMe
4
.[note 1][13]

This large number of stable isotopes is thought to be a direct result of tin possessing an atomic number of 50, which is a "magic number" in nuclear physics. In addition, there are 29 known unstable isotopes, encompassing all the remaining ones with atomic masses between 99 and 137. Aside from 126Sn, which has a half-life of 230,000 years, all the radioactive isotopes have a half-life of less than a year. The radioactive 100Sn is one of the few nuclides possessing a "doubly magic" nucleus and was discovered in 1994.[14] Another 30 metastable isomers have been characterized for isotopes between 111 and 131, the most stable being 121mSn, with a half-life of 43.9 years.

Etymology[edit]

The word tin is shared among Germanic languages and can be traced back to reconstructed Proto-Germanic *tin-om; cognates include German Zinn, Swedish tenn and Dutch tin. It is not found in other branches of Indo-European, except by borrowing from Germanic (e.g. Irish tinne from English).[15][16]

The Latin name stannum originally meant an alloy of silver and lead, and came to mean 'tin' in the 4th century BCE[17]—the earlier Latin word for it was plumbum candidum, or "white lead". Stannum apparently came from an earlier stāgnum (meaning the same substance),[15] the origin of the Romance and Celtic terms for 'tin'.[15][18] The origin of stannum/stāgnum is unknown; it may be pre-Indo-European.[19] The Meyers Konversationslexikon speculates on the contrary that stannum is derived from (the ancestor of) Cornish stean, and is proof that Cornwall in the first centuries AD was the main source of tin.

History[edit]

Ceremonial giant bronze dirk of the Plougrescant-Ommerschans type, Plougrescant, France, 1500–1300 BC.

Tin extraction and use can be dated to the beginnings of the Bronze Age around 3000 BC, when it was observed that copper objects formed of polymetallic ores with different metal contents had different physical properties.[20] The earliest bronze objects had a tin or arsenic content of less than 2% and are therefore believed to be the result of unintentional alloying due to trace metal content in the copper ore.[21] The addition of a second metal to copper increases its hardness, lowers the melting temperature, and improves the casting process by producing a more fluid melt that cools to a denser, less spongy metal.[21] This was an important innovation that allowed for the much more complex shapes cast in closed moulds of the Bronze Age. Arsenical bronze objects appear first in the Near East where arsenic is commonly found in association with copper ore, but the health risks were quickly realized and the quest for sources of the much less hazardous tin ores began early in the Bronze Age.[22] This created the demand for rare tin metal and formed a trade network that linked the distant sources of tin to the markets of Bronze Age cultures.[citation needed]

Cassiterite (SnO2), the tin oxide form of tin, was most likely the original source of tin in ancient times. Other forms of tin ores are less abundant sulfides such as stannite that require a more involved smelting process. Cassiterite often accumulates in alluvial channels as placer deposits due to the fact that it is harder, heavier, and more chemically resistant than the granite in which it typically forms.[23] These deposits can be easily seen in river banks as cassiterite is usually black, purple or otherwise dark in color, a feature exploited by early Bronze Age prospectors. It is likely that the earliest deposits were alluvial in nature, and perhaps exploited by the same methods used for panning gold in placer deposits.[citation needed]

Compounds and chemistry[edit]

In the great majority of its compounds, tin has the oxidation state II or IV.

Inorganic compounds[edit]

Halide compounds are known for both oxidation states. For Sn(IV), all four halides are well known: SnF4, SnCl4, SnBr4, and SnI4. The three heavier members are volatile molecular compounds, whereas the tetrafluoride is polymeric. All four halides are known for Sn(II) also: SnF2, SnCl2, SnBr2, and SnI2. All are polymeric solids. Of these eight compounds, only the iodides are colored.[24]

Tin(II) chloride (also known as stannous chloride) is the most important tin halide in a commercial sense. Illustrating the routes to such compounds, chlorine reacts with tin metal to give SnCl4 whereas the reaction of hydrochloric acid and tin gives SnCl2 and hydrogen gas. Alternatively SnCl4 and Sn combine to stannous chloride via a process called comproportionation:[25]

SnCl4 + Sn → 2 SnCl2

Tin can form many oxides, sulfides, and other chalcogenide derivatives. The dioxide SnO2 (cassiterite) forms when tin is heated in the presence of air.[24] SnO2 is amphoteric, which means that it dissolves in both acidic and basic solutions.[26] There are also stannates with the structure [Sn(OH)6]2−, like K2[Sn(OH)6], although the free stannic acid H2[Sn(OH)6] is unknown. The sulfides of tin exist in both the +2 and +4 oxidation states: tin(II) sulfide and tin(IV) sulfide (mosaic gold).

Ball-and-stick models of the structure of solid stannous chloride (SnCl2).[27]

Hydrides[edit]

Stannane (SnH4), where tin is in the +4 oxidation state, is unstable. Organotin hydrides are however well known, e.g. tributyltin hydride (Sn(C4H9)3H).[5] These compound release transient tributyl tin radicals, rare examples of compounds of tin(III).[28]

Organotin compounds[edit]

Organotin compounds, sometimes called stannanes, are chemical compounds with tin–carbon bonds.[29] Of the compounds of tin, the organic derivatives are the most useful commercially.[30] Some organotin compounds are highly toxic and have been used as biocides. The first organotin compound to be reported was diethyltin diiodide ((C2H5)2SnI2), reported by Edward Frankland in 1849.[31]

Most organotin compounds are colorless liquids or solids that are stable to air and water. They adopt tetrahedral geometry. Tetraalkyl- and tetraaryltin compounds can be prepared using Grignard reagents:[30]

SnCl
4
+ 4 RMgBr → R
4
Sn
+ 4 MgBrCl

The mixed halide-alkyls, which are more common and more important commercially than the tetraorgano derivatives, are prepared by redistribution reactions:

SnCl
4
+ R
4
Sn
→ 2 SnCl2R2

Divalent organotin compounds are uncommon, although more common than related divalent organogermanium and organosilicon compounds. The greater stabilization enjoyed by Sn(II) is attributed to the "inert pair effect". Organotin(II) compounds include both stannylenes (formula: R2Sn, as seen for singlet carbenes) and distannylenes (R4Sn2), which are roughly equivalent to alkenes. Both classes exhibit unusual reactions.[32]

Occurrence[edit]

Sample of cassiterite, the main ore of tin.
Granular pieces of cassiterite, which are collected by placer mining

Tin is generated via the long S-process in low-to-medium mass stars (with masses of 0.6 to 10 times that of Sun). It arises via beta decay of heavy isotopes of indium.[33]

Tin is the 49th most abundant element in Earth's crust, representing 2 ppm compared with 75 ppm for zinc, 50 ppm for copper, and 14 ppm for lead.[34]

Tin does not occur as the native element but must be extracted from various ores. Cassiterite (SnO2) is the only commercially important source of tin, although small quantities of tin are recovered from complex sulfides such as stannite, cylindrite, franckeite, canfieldite, and teallite. Minerals with tin are almost always associated with granite rock, usually at a level of 1% tin oxide content.[35]

Because of the higher specific gravity of tin dioxide, about 80% of mined tin is from secondary deposits found downstream from the primary lodes. Tin is often recovered from granules washed downstream in the past and deposited in valleys or under sea. The most economical ways of mining tin are through dredging, hydraulic methods or open cast mining. Most of the world's tin is produced from placer deposits, which may contain as little as 0.015% tin.[36]

World tin mine reserves (tonnes, 2011)[37]
Country Reserves
 China 1,500,000
 Malaysia 250,000
 Peru 310,000
 Indonesia 800,000
 Brazil 590,000
 Bolivia 400,000
 Russia 350,000
 Australia 180,000
 Thailand 170,000
  Other 180,000
  Total 4,800,000

About 253,000 tonnes of tin have been mined in 2011, mostly in China (110,000 t), Indonesia (51,000 t), Peru (34,600 t), Bolivia (20,700 t) and Brazil (12,000 t).[37] Estimates of tin production have historically varied with the dynamics of economic feasibility and the development of mining technologies, but it is estimated that, at current consumption rates and technologies, the Earth will run out of tin that can be mined in 40 years.[38] However Lester Brown has suggested tin could run out within 20 years based on an extremely conservative extrapolation of 2% growth per year.[39]

Economically recoverable tin reserves[35]
Year Million tonnes
1965 4,265
1970 3,930
1975 9,060
1980 9,100
1985 3,060
1990 7,100
2000 7,100[37]
2010 5,200[37]

Secondary, or scrap, tin is also an important source of the metal. The recovery of tin through secondary production, or recycling of scrap tin, is increasing rapidly. Whereas the United States has neither mined since 1993 nor smelted tin since 1989, it was the largest secondary producer, recycling nearly 14,000 tonnes in 2006.[37]

New deposits are reported to be in southern Mongolia,[40] and in 2009, new deposits of tin were discovered in Colombia, by the Seminole Group Colombia CI, SAS.[41]

Production[edit]

Tin is produced by carbothermic reduction of the oxide ore with carbon or coke. Both reverberatory furnace and electric furnace can be used.[42][43][44]

Mining and smelting[edit]

Main article: Tin mining

Industry[edit]

Candlestick made of tin

The ten largest companies produced most of the world's tin in 2007. It is not clear which of these companies include tin smelted from the mine at Bisie, Democratic Republic of the Congo, which is controlled by a renegade militia and produces 15,000 tonnes. Most of the world's tin is traded on the London Metal Exchange (LME), from 8 countries, under 17 brands.[45]

Largest tin producing companies (tonnes)[46]
Company Polity 2006 2007 %Change
Yunnan Tin China 52,339 61,129 16.7
PT Timah Indonesia 44,689 58,325 30.5
Minsur Peru 40,977 35,940 −12.3
Malay China 52,339 61,129 16.7
Malaysia Smelting Corp Malaysia 22,850 25,471 11.5
Thaisarco Thailand 27,828 19,826 −28.8
Yunnan Chengfeng China 21,765 18,000 −17.8
Liuzhou China Tin China 13,499 13,193 −2.3
EM Vinto Bolivia 11,804 9,448 −20.0
Gold Bell Group China 4,696 8,000 70.9

Price and exchanges[edit]

World production and price (US exchange) of tin.

Tin is unique among other mineral commodities by the complex "agreements" between producer countries and consumer countries dating back to 1921. The earlier agreements tended to be somewhat informal and sporadic; they led to the "First International Tin Agreement" in 1956, the first of a continuously numbered series that essentially collapsed in 1985. Through this series of agreements, the International Tin Council (ITC) had a considerable effect on tin prices. The ITC supported the price of tin during periods of low prices by buying tin for its buffer stockpile and was able to restrain the price during periods of high prices by selling tin from the stockpile. This was an anti-free-market approach, designed to assure a sufficient flow of tin to consumer countries and a decent profit for producer countries. However, the buffer stockpile was not sufficiently large, and during most of those 29 years tin prices rose, sometimes sharply, especially from 1973 through 1980 when rampant inflation plagued many world economies.[47]

During the late 1970s and early 1980s, the U.S. Government tin stockpile was in an aggressive selling mode, partly to take advantage of the historically high tin prices. The sharp recession of 1981–82 proved to be quite harsh on the tin industry. Tin consumption declined dramatically. The ITC was able to avoid truly steep declines through accelerated buying for its buffer stockpile; this activity required the ITC to borrow extensively from banks and metal trading firms to augment its resources. The ITC continued to borrow until late 1985, when it reached its credit limit. Immediately, a major "tin crisis" followed — tin was delisted from trading on the London Metal Exchange for about three years, the ITC dissolved soon afterward, and the price of tin, now in a free-market environment, plummeted sharply to $4 per pound and remained around this level through the 1990s.[47] It increased again by 2010 due to the rebound in consumption following the 2008–09 world economic crisis, restocking and continued growth in consumption by the world's developing economies.[37]

London Metal Exchange (LME) is the principal trading site for tin.[37] Other tin contract markets are Kuala Lumpur Tin Market (KLTM) and Indonesia Tin Exchange (INATIN).[48]

Price of tin in USD cents per kg[edit]

Tin (USD cents per kg)
2008 2009 2010 2011 2012
Price 1,851 1,357 2,041 2,605 2,113

Source:Helgi Library [49]

Applications[edit]

World consumption of refined tin by end use, 2006

In 2006, about half of tin produced was used in solder. The rest was divided between tin plating, tin chemicals, brass and bronze, and niche uses.[50]

Solder[edit]

A coil of lead-free solder wire

Tin has long been used as a solder in the form of an alloy with lead, tin accounting for 5 to 70% w/w. Tin forms a eutectic mixture with lead containing 63% tin and 37% lead. Such solders are primarily used for joining pipes or electric circuits. Since the European Union Waste Electrical and Electronic Equipment Directive (WEEE Directive) and Restriction of Hazardous Substances Directive came into effect on 1 July 2006, the use of lead in such alloys has decreased. Replacing lead has many problems, including a higher melting point, and the formation of tin whiskers causing electrical problems. Tin pest can occur in lead-free solders, leading to loss of the soldered joint. Replacement alloys are rapidly being found, although problems of joint integrity remain.[51]

Tin plating[edit]

Tin bonds readily to iron and is used for coating lead, zinc and steel to prevent corrosion. Tin-plated steel containers are widely used for food preservation, and this forms a large part of the market for metallic tin. A tinplate canister for preserving food was first manufactured in London in 1812.[52] Speakers of British English call them "tins", while speakers of American English call them "cans" or "tin cans". One thus-derived use of the slang term "tinnie" or "tinny" means "can of beer". The tin whistle is so called because it was first mass-produced in tin-plated steel.[53][54]

Specialized alloys[edit]

Pewter plate

Tin in combination with other elements forms a wide variety of useful alloys. Tin is most commonly alloyed with copper. Pewter is 85–99% tin;[55] Bearing metal has a high percentage of tin as well.[56][57] Bronze is mostly copper (12% tin), while addition of phosphorus gives phosphor bronze. Bell metal is also a copper-tin alloy, containing 22% tin. Tin has also sometimes been used in coinage; for example, it once formed a single-digit figure percentage (usually five percent or less) of the American[58] and Canadian[59] pennies. Because copper is often the major metal in such coins, and zinc is sometimes present as well, these could technically be called bronze and/or brass alloys.

Tin plated metal from a can
Artisan Alfonso Santiago Leyva and his son working tin sheets

The niobium-tin compound Nb3Sn is commercially used as wires for superconducting magnets, due to the material's high critical temperature (18 K) and critical magnetic field (25 T). A superconducting magnet weighing as little as two kilograms is capable of producing magnetic fields comparable to a conventional electromagnet weighing tons.[60]

The addition of a few percent of tin is commonly used in zirconium alloys for the cladding of nuclear fuel.[61]

Most metal pipes in a pipe organ are made of varying amounts of a tin/lead alloy, with 50%/50% being the most common. The amount of tin in the pipe defines the pipe's tone, since tin is the most tonally resonant of all metals.[dubious ] When a tin/lead alloy cools, the lead cools slightly faster and produces a mottled or spotted effect. This metal alloy is referred to as spotted metal. Major advantages of using tin for pipes include its appearance, its workability, and resistance to corrosion.[62][63]

Other applications[edit]

A 21st-century reproduction barn lantern made of punched tin.

Punched tin-plated steel, also called pierced tin, is an artisan technique originating in central Europe for creating housewares that are both functional and decorative. Decorative piercing designs exist in a wide variety, based on geography or the artisan's personal creations. Punched tin lanterns are the most common application of this artisan technique. The light of a candle shining through the pierced design creates a decorative light pattern in the room where it sits. Punched tin lanterns and other punched tin articles were created in the New World from the earliest European settlement. A well-known example is the Revere type lantern, named after Paul Revere.[64]

Before the modern era, in some areas of the Alps, a goat or sheep's horn would be sharpened and a tin panel would be punched out using the alphabet and numbers from one to nine. This learning tool was known appropriately as "the horn". Modern reproductions are decorated with such motifs as hearts and tulips.

In America, pie safes and food safes came into use in the days before refrigeration. These were wooden cupboards of various styles and sizes – either floor standing or hanging cupboards meant to discourage vermin and insects and to keep dust from perishable foodstuffs. These cabinets had tinplate inserts in the doors and sometimes in the sides, punched out by the homeowner, cabinetmaker or a tinsmith in varying designs to allow for air circulation. Modern reproductions of these articles remain popular in North America.[65]

Window glass is most often made by floating molten glass on top of molten tin (creating float glass) in order to produce a flat surface. This is called the "Pilkington process".[66]

Tin is also used as a negative electrode in advanced Li-ion batteries. Its application is somewhat limited by the fact that some tin surfaces[which?] catalyze decomposition of carbonate-based electrolytes used in Li-ion batteries.[67]

Tin(II) fluoride is added to some dental care products[68] as stannous fluoride (SnF2). Tin(II) fluoride can be mixed with calcium abrasives while the more common sodium fluoride gradually becomes biologically inactive combined with calcium compounds.[69] It has also been shown to be more effective than sodium fluoride in controlling gingivitis.[70]

Organotin compounds[edit]

Main article: Organotin chemistry

Of all the chemical compounds of tin, the organotin compounds are most heavily used. Worldwide industrial production probably exceeds 50,000 tonnes.[71]

PVC stabilizers[edit]

The major commercial application of organotin compounds is in the stabilization of PVC plastics. In the absence of such stabilizers, PVC would otherwise rapidly degrade under heat, light, and atmospheric oxygen, to give discolored, brittle products. Tin scavenges labile chloride ions (Cl), which would otherwise initiate loss of HCl from the plastic material.[72] Typical tin compounds are carboxylic acid derivatives of dibutyltin dichloride, such as the dilaurate.[73]

Biocides[edit]

Organotin compounds can have a relatively high toxicity, which is both advantageous and problematic. They have been used for their biocidal effects in/as fungicides, pesticides, algaecides, wood preservatives, and antifouling agents.[72] Tributyltin oxide is used as a wood preservative.[74] Tributyltin was used as additive for ship paint to prevent growth of marine organisms on ships, with use declining after organotin compounds were recognized as persistent organic pollutants with an extremely high toxicity for some marine organisms, for example the dog whelk.[75] The EU banned the use of organotin compounds in 2003,[76] while concerns over the toxicity of these compounds to marine life and their effects on the reproduction and growth of some marine species,[72] (some reports describe biological effects to marine life at a concentration of 1 nanogram per liter) have led to a worldwide ban by the International Maritime Organization.[77] Many nations now restrict the use of organotin compounds to vessels over 25 meters long.[72]

Organic chemistry[edit]

Some tin reagents are useful in organic chemistry. In the largest application, stannous chloride is a common reducing agent for the conversion of nitro and oxime groups to amines. The Stille reaction couples organotin compounds with organic halides or pseudohalides.[78]

Li-ion batteries[edit]

Tin forms several inter-metallic phases with lithium metal and it makes it a potentially attractive material. Large volumetric expansion of tin upon alloying with lithium and instability of the tin-organic electrolyte interface at low electrochemical potentials are the greatest challenges in employing it in commercial cells. The problem was partially solved by Sony. Tin inter-metallic compound with cobalt, mixed with carbon, has been implemented by Sony in its Nexelion cells released in late 2000's. The composition of the active materials is close to Sn0.3Co0.4C0.3. Recent research showed that only some crystalline facets of tetragonal (beta) Sn are responsible for undesirable electrochemical activity.[79]

Precautions[edit]

Main article: Tin poisoning

Cases of poisoning from tin metal, its oxides, and its salts are "almost unknown". On the other hand, certain organotin compounds are almost as toxic as cyanide.[30]

People can be exposed to tin in the workplace by breathing it in, skin contact, and eye contact. The Occupational Safety and Health Administration (OSHA) has set the legal limit (Permissible_exposure_limit) for tin exposure in the workplace as 2 mg/m3 over an 8-hour workday. The National_Institute_for_Occupational_Safety_and_Health (NIOSH) has set a recommended exposure limit (REL) of 2 mg/m3 over an 8-hour workday. At levels of 100 mg/m3, tin is immediately dangerous to life and health.[80]

See also[edit]

Notes[edit]

  1. ^ Only H, F, P, Tl and Xe have a higher receptivity for NMR analysis for samples containing isotopes at their natural abundance.

References[edit]

  1. ^ Standard Atomic Weights 2013. Commission on Isotopic Abundances and Atomic Weights
  2. ^ "SnH3". NIST Chemistry WebBook. National Institure of Standards and Technology. Retrieved 23 January 2013. 
  3. ^ "HSn". NIST Chemistry WebBook. National Institute of Standards and Technology. Retrieved 23 January 2013. 
  4. ^ Lide, D. R., ed. (2005). "Magnetic susceptibility of the elements and inorganic compounds". CRC Handbook of Chemistry and Physics (PDF) (86th ed.). Boca Raton (FL): CRC Press. ISBN 0-8493-0486-5. 
  5. ^ a b c d e Holleman, Arnold F.; Wiberg, Egon; Wiberg, Nils (1985). "Tin". Lehrbuch der Anorganischen Chemie (in German) (91–100 ed.). Walter de Gruyter. pp. 793–800. ISBN 3-11-007511-3. 
  6. ^ Ink with tin nanoparticles could print future circuit boards, Physorg, April 12, 2011; Jo, Yun Hwan; Jung, Inyu; Choi, Chung Seok; Kim, Inyoung; Lee, Hyuck Mo (2011). "Synthesis and characterization of low temperature Sn nanoparticles for the fabrication of highly conductive ink". Nanotechnology 22 (22): 225701. Bibcode:2011Nanot..22v5701J. doi:10.1088/0957-4484/22/22/225701. PMID 21454937. 
  7. ^ Molodets, A. M.; Nabatov, S. S. (2000). "Thermodynamic Potentials, Diagram of State, and Phase Transitions of Tin on Shock Compression". High Temperature 38 (5): 715–721. doi:10.1007/BF02755923. 
  8. ^ This conversion is known as tin disease or tin pest. Tin pest was a particular problem in northern Europe in the 18th century as organ pipes made of tin alloy would sometimes be affected during long cold winters. Some unverifiable sources also say that, during Napoleon's Russian campaign of 1812, the temperatures became so cold that the tin buttons on the soldiers' uniforms disintegrated over time, contributing to the defeat of the Grande Armée.Le Coureur, Penny; Burreson, Jay (2004). Napoleon's Buttons: 17 Molecules that Changed History. New York: Penguin Group USA. , a persistent legend that probably has no background in real events. Öhrström, Lars (2013). The Last Alchemist in Paris. Oxford: Oxford University Press. ISBN 978-0-19-966109-1.  Cotton, Simon (2014). "Book review: The last alchemist in Paris". Chemistry World. http://rsc.li/CW_140501
  9. ^ a b Schwartz, Mel (2002). "Tin and Alloys, Properties". Encyclopedia of Materials, Parts and Finishes (2nd ed.). CRC Press. ISBN 1-56676-661-3. 
  10. ^ Dehaas, W; Deboer, J; Vandenberg, G (1935). "The electrical resistance of cadmium, thallium and tin at low temperatures". Physica 2: 453. Bibcode:1935Phy.....2..453D. doi:10.1016/S0031-8914(35)90114-8. 
  11. ^ Meissner, W.; R. Ochsenfeld (1933). "Ein neuer effekt bei eintritt der Supraleitfähigkeit". Naturwissenschaften 21 (44): 787–788. Bibcode:1933NW.....21..787M. doi:10.1007/BF01504252. 
  12. ^ Craig, Bruce D; Anderson, David S; International, A.S.M. (January 1995). Handbook of corrosion data. p. 126. ISBN 978-0-87170-518-1. 
  13. ^ "Interactive NMR Frequency Map". Archived from the original on 2011-06-04. Retrieved 2009-05-05. 
  14. ^ Walker, Phil (1994). "Doubly Magic Discovery of Tin-100". Physics World 7 (June). 
  15. ^ a b c "tin". Oxford English Dictionary (3rd ed.). Oxford University Press. September 2005.  (Subscription or UK public library membership required.)
  16. ^ Harper, Douglas. "tin". Online Etymology Dictionary. 
  17. ^ Encyclopædia Britannica, 11th Edition, 1911, s.v. 'tin', citing H. Kopp
  18. ^ "The Ancient Mining of Tin". oxleigh.freeserve.co.uk. Archived from the original on 2009-04-03. Retrieved 2009-07-07. 
  19. ^ American Heritage Dictionary
  20. ^ Cierny, J.; Weisgerber, G. (2003). "The "Bronze Age tin mines in Central Asia". In Giumlia-Mair, A.; Lo Schiavo, F. The Problem of Early Tin. Oxford: Archaeopress. pp. 23–31. ISBN 1-84171-564-6. 
  21. ^ a b Penhallurick, R.D. (1986). Tin in Antiquity: its Mining and Trade Throughout the Ancient World with Particular Reference to Cornwall. London: The Institute of Metals. ISBN 0-904357-81-3. 
  22. ^ Charles, J.A. (1979). "The development of the usage of tin and tin-bronze: some problems". In Franklin, A.D.; Olin, J.S.; Wertime, T.A. The Search for Ancient Tin. Washington D.C.: A seminar organized by Theodore A. Wertime and held at the Smithsonian Institution and the National Bureau of Standards, Washington D.C. March 14–15, 1977. pp. 25–32. 
  23. ^ (Penhallurick 1986)
  24. ^ a b Holleman, A. F.; Wiberg, E. (2001), Inorganic Chemistry, San Diego: Academic Press, ISBN 0-12-352651-5 
  25. ^ Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. ISBN 0080379419. 
  26. ^ Inorganic & Theoretical chemistry, F. Sherwood Taylor, Heineman, 6th Edition (1942)
  27. ^ J. M. Leger; J. Haines; A. Atouf (1996). "The high pressure behaviour of the cotunnite and post-cotunnite phases of PbCl2 and SnCl2". J. Phys. Chem. Solids 57 (1): 7–16. Bibcode:1996JPCS...57....7L. doi:10.1016/0022-3697(95)00060-7. 
  28. ^ Gaur, D. P.; Srivastava, G.; Mehrotra, R. C. (1973). "Organic Derivatives of Tin. III. Reactions of Trialkyltin Ethoxide with Alkanolamines". Zeitschrift für anorganische und allgemeine Chemie 398: 72. doi:10.1002/zaac.19733980109. 
  29. ^ Elschenbroich, C. "Organometallics" (2006) Wiley-VCH: Weinheim. ISBN 978-3-527-29390-2
  30. ^ a b c Graf, G. G. (2000) "Tin, Tin Alloys, and Tin Compounds" in Ullmann's Encyclopedia of Industrial Chemistry, 2005 Wiley-VCH, Weinheim doi:10.1002/14356007.a27_049
  31. ^ Sander H.L. Thoonen; Berth-Jan Deelman; Gerard van Koten (2004). "Synthetic aspects of tetraorganotins and organotin(IV) halides" (PDF). Journal of Organometallic Chemistry (689): 2145–2157. 
  32. ^ Peng, Yang; Ellis, Bobby D.; Wang, Xinping; Fettinger, James C.; Power, P. P. (2009). "Reversible Reactions of Ethylene with Distannynes Under Ambient Conditions". Science 325 (5948): 1668–1670. Bibcode:2009Sci...325.1668P. doi:10.1126/science.1176443. PMID 19779193. 
  33. ^ Shu, Frank H (1982). "The physical universe: An introduction to astronomy". pp. 119–121. ISBN 978-0-935702-05-7. 
  34. ^ Emsley 2001, pp. 124, 231, 449 and 503
  35. ^ a b "Tin: From Ore to Ingot". International Tin Research Institute. 1991. Archived from the original on 2009-03-22. Retrieved 2009-03-21. 
  36. ^ Sutphin, David M; Reed, David M. Sutphin Andrew E. Sabin Bruce L; Sabin, Andrew E; Reed, Bruce L (1992-06-01). Tin – International Strategic Minerals Inventory Summary Report. p. 9. ISBN 978-0-941375-62-7. 
  37. ^ a b c d e f g Carlin, Jr., James F. "Tin: Statistics and Information" (PDF). United States Geological Survey. Retrieved 2008-11-23. 
  38. ^ Reilly, Michael (May 26, 2007). "How Long Will it Last?". New Scientist 194 (2605): 38–39. Bibcode:2007NewSc.194...38R. doi:10.1016/S0262-4079(07)61508-5. ISSN 0262-4079. 
  39. ^ Brown, Lester (2006). Plan B 2.0. New York: W.W. Norton. p. 109. ISBN 978-0-393-32831-8. 
  40. ^ Kovalenko, V. I.; Yarmolyuk, V. V. (1995). "Endogenous rare metal ore formations and rare metal metallogeny of Mongolia". Economic Geology 90 (3): 520. doi:10.2113/gsecongeo.90.3.520. 
  41. ^ "Seminole Group Colombia Discovers High Grade Tin Ore in the Amazon Jungle". FreePR101 PressRelease. 
  42. ^ Schrader, George F; Elshennawy, Ahmad K; Doyle, Lawrence E (July 2000). Manufacturing processes and materials. ISBN 978-0-87263-517-3. 
  43. ^ Louis, Henry (1911). Metallurgy of tin. McGraw-Hill book Company. 
  44. ^ Knorr, Klaus (1945). Tin Under Control. Stanford University Press. ISBN 978-0-8047-2136-3. 
  45. ^ "International Tin Research Institute. LME Tin Brands". ITRI. Archived from the original on 2008-12-07. Retrieved 2009-05-05. 
  46. ^ "International Tin Research Institute. Top Ten Tin Producing Companies.". Archived from the original on 2008-12-07. Retrieved 2009-05-05. 
  47. ^ a b Carlin, James F., Jr. (1998). Significant events affecting tin prices since 1958. USGS.
  48. ^ "12 Januari Pemasaran Perdana INATIN". December 15, 2011. 
  49. ^ | http://helgilibrary.com/indicators/index/tin-usd-cents-per-kg | 2014-02-10
  50. ^ "Tin Use Survey 2007". ITRI. Archived from the original on 2008-12-07. Retrieved 2008-11-21. 
  51. ^ Black, Harvey (2005). "Getting the Lead Out of Electronics". Environmental Health Perspectives 113 (10): A682–5. doi:10.1289/ehp.113-a682. PMC 1281311. PMID 16203230. 
  52. ^ Education in chemistry 32: 92–. 1995.  Missing or empty |title= (help)
  53. ^ Control, Tin Under (1945). Tin Under Control. pp. 10–15. ISBN 978-0-8047-2136-3. 
  54. ^ Panel On Tin, National Research Council (U.S.). Committee on Technical Aspects of Critical and Strategic Materials (1970). Trends in the use of tin. pp. 10–22. 
  55. ^ Hull, Charles (1992). Pewter. Osprey Publishing. pp. 1–5. ISBN 978-0-7478-0152-8. 
  56. ^ Brakes, James (2009). "Introduction". Analysis of Babbit. BiblioBazaar, LLC. pp. 1–2. ISBN 978-1-110-11092-6. 
  57. ^ Williams, Robert S. (2007). Principles of Metallography. Read books. pp. 46–47. ISBN 978-1-4067-4671-6. 
  58. ^ "The Composition of the Cent". US Mint. Retrieved 2011-10-28. 
  59. ^ "Composition of canadian coins" (PDF). Canadian Mint. Retrieved 2011-10-28. 
  60. ^ Geballe, Theodore H. (October 1993). "Superconductivity: From Physics to Technology". Physics Today 46 (10): 52–56. Bibcode:1993PhT....46j..52G. doi:10.1063/1.881384. 
  61. ^ Campbell, Flake C (2008). "Zirconium". Elements of Metallurgy and Engineering Alloys. p. 597. ISBN 978-0-87170-867-0. 
  62. ^ Robert Palmieri, ed. (2006). "Pipe Metal". Encyclopedia of keyboard instruments. New York: Garland. p. 411. ISBN 978-0-415-94174-7. 
  63. ^ George Ashdown Audsley (1988). "Metal Pipes: And the Materials used in their Construction". The Art of Organ Building Audsley, George Ashdown. Courier Dover Publications. p. 501. ISBN 978-0-486-21315-6. 
  64. ^ Bridge, Janet (September 1996). Making & decorating picture frames. ISBN 978-0-89134-739-2. 
  65. ^ "Tin punching". Retrieved August 15, 2011. 
  66. ^ Pilkington, L. A. B. (1969). "Review Lecture. The Float Glass Process.". Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences 314 (1516): 1–25. Bibcode:1969RSPSA.314....1P. doi:10.1098/rspa.1969.0212. JSTOR 2416528. 
  67. ^ Lucas, Ivan T.; Syzdek, Jarosław; Kostecki, Robert (2011). "Interfacial processes at single-crystal β-Sn electrodes in organic carbonate electrolytes". Electrochemistry Communications 13 (11): 1271–1275. doi:10.1016/j.elecom.2011.08.026. 
  68. ^ "Colgate Gel-Kam". Retrieved 2009-05-05. 
  69. ^ Hattab, F. (April 1989). "The State of Fluorides in Toothpastes". Journal of Dentistry 17 (2): 47–54. doi:10.1016/0300-5712(89)90129-2. PMID 2732364. 
  70. ^ Perlich, MA; Bacca, LA; Bollmer, BW; Lanzalaco, AC; McClanahan, SF; Sewak, LK; Beiswanger, BB; Eichold, WA; Hull, JR; et al. (1995). "The clinical effect of a stabilized stannous fluoride dentifrice on plaque formation, gingivitis and gingival bleeding: a six-month study". The Journal of Clinical Dentistry 6 (Special Issue): 54–58. PMID 8593194. 
  71. ^ Ebdon, L; Britain), Royal Society of Chemistry (Great (2001). "Organotin in Industrial and Domestic Products". Trace element speciation for environment, food and health. p. 144. ISBN 978-0-85404-459-7. 
  72. ^ a b c d Atkins, Peter; Shriver, Duward F.; Overton, Tina & Rourke, Jonathan (2006). Inorganic chemistry (4th ed.). W.H. Freeman. pp. 343, 345. ISBN 0-7167-4878-9. 
  73. ^ Wilkes, Charles E; Summers, James W; Daniels, Charles Anthony; Berard, Mark T (August 2005). PVC handbook. p. 108. ISBN 978-1-56990-379-7. 
  74. ^ David N.-S. Hon; Nobuo Shiraishi, eds. (2001). "Preservation of Wood". Wood and cellulosic chemistry. New York, NY: Dekker. p. 799. ISBN 978-0-8247-0024-9. 
  75. ^ Eisler, Ronald. "Tin Hazards To Fish, Wildlife, and Invertebrates: A Synoptic Review" (PDF). U.S. Fish and Wildlife Service Patuxent Wildlife Research Center. 
  76. ^ "Regulation (EC) No 782/2003 of the European Parliament and of the Council of 14 April 2003 on the prohibition of organotin compounds on ships". Retrieved 2009-05-05. 
  77. ^ Dürr, Simone; Thomason, Jeremy, eds. (2008). "Fouling on Shipping". Biofouling. Oxford: Blackwell. p. 227. ISBN 978-1-4051-6926-4. 
  78. ^ Farina, Vittorio; Krishnamurthy, Venkat; Scott, William J. (1997). "The Stille Reaction". Organic Reactions (New York: Wiley). doi:10.1002/0471264180.or050.01. ISBN 0-471-31273-8. 
  79. ^ Lucas, Ivan; Syzdek, Jaroslaw (2011). "Electrochemistry Communications". Electrochemistry Communications 13 (11): 1271. doi:10.1016/j.elecom.2011.08.026. 
  80. ^ "CDC - NIOSH Pocket Guide to Chemical Hazards - Tin". www.cdc.gov. Retrieved 2015-11-24. 

Bibliography[edit]

External links[edit]