Jump to content

Sulfur: Difference between revisions

From Wikipedia, the free encyclopedia
[pending revision][pending revision]
Content deleted Content added
m Reverted edits by Prettypolitico (talk) to last version by N1ywb
Judygt (talk | contribs)
Line 243: Line 243:
*[http://www.sulphurinstitute.org/ The Sulphur Institute]
*[http://www.sulphurinstitute.org/ The Sulphur Institute]
*''[http://www.tufts.edu/as/wright_center/cosmic_evolution/docs/text/text_stel_6.html Origin of Heavy Elements,]'' by Tufts University {{Clear}}
*''[http://www.tufts.edu/as/wright_center/cosmic_evolution/docs/text/text_stel_6.html Origin of Heavy Elements,]'' by Tufts University {{Clear}}
* [http://www.farmchemicalsinternational.com/cropprotection/cpd/?op=cpdproductdetail&pid=368100 Crop Protection Database: Learn more about sulfur]


{{Compact periodic table}}
{{Compact periodic table}}

Revision as of 13:37, 7 April 2011

Sulfur, 16S
Sulfur
Alternative nameSulphur (British spelling)
Allotropessee Allotropes of sulfur
AppearanceLemon yellow sintered microcrystals
Standard atomic weight Ar°(S)
Sulfur 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
O

S

Se
phosphorussulfurchlorine
Atomic number (Z)16
Groupgroup 16 (chalcogens)
Periodperiod 3
Block  p-block
Electron configuration[Ne] 3s2 3p4
Electrons per shell2, 8, 6
Physical properties
Phase at STPsolid
Melting pointalpha (α-S8): 388.36 K ​(115.21 °C, ​239.38 °F)
Boiling point717.8 K ​(444.6 °C, ​832.3 °F)
Density (near r.t.)alpha (α-S8): 2.07 g/cm3
beta (β-S8): 1.96 g/cm3
gamma (γ-S8): 1.92 g/cm3
when liquid (at m.p.)1.819 g/cm3
Critical point1314 K, 20.7 MPa
Heat of fusionbeta (β-S8): 1.727 kJ/mol
Heat of vaporizationbeta (β-S8): 45 kJ/mol
Molar heat capacity22.75 J/(mol·K)
Vapor pressure
P (Pa) 1 10 100 1 k 10 k 100 k
at T (K) 375 408 449 508 591 717
Atomic properties
Oxidation states−2, −1, 0, +1, +2, +3, +4, +5, +6 (a strongly acidic oxide)
ElectronegativityPauling scale: 2.58
Ionization energies
  • 1st: 999.6 kJ/mol
  • 2nd: 2252 kJ/mol
  • 3rd: 3357 kJ/mol
  • (more)
Covalent radius105±3 pm
Van der Waals radius180 pm
Color lines in a spectral range
Spectral lines of sulfur
Other properties
Natural occurrenceprimordial
Crystal structurealpha (α-S8): ​orthorhombic (oF128)
Lattice constants
Orthorhombic crystal structure for alpha (α-S8): sulfur
a = 1.0460 nm
b = 1.2861 nm
c = 2.4481 nm (at 20 °C)[3]
Crystal structurebeta (β-S8): ​monoclinic (mP48)
Lattice constants
Monoclinic crystal structure for beta (β-S8): sulfur
a = 1.0923 nm
b = 1.0851 nm
c = 1.0787 nm
β = 95.905° (at 20 °C)[3]
Thermal conductivity0.205 W/(m⋅K) (amorphous)
Electrical resistivity2×1015  Ω⋅m (at 20 °C) (amorphous)
Magnetic orderingdiamagnetic[4]
Molar magnetic susceptibilityalpha (α-S8): −15.5×10−6 cm3/mol (298 K)[5]
Bulk modulus7.7 GPa
Mohs hardness2.0
CAS Number7704-34-9
History
Discoverybefore 2000 BCE[6]
Recognized as an element byAntoine Lavoisier (1777)
Isotopes of sulfur
Main isotopes Decay
abun­dance half-life (t1/2) mode pro­duct
32S 94.8% stable
33S 0.760% stable
34S 4.37% stable
35S trace 87.37 d β 35Cl
36S 0.02% stable
34S abundances vary greatly (between 3.96 and 4.77 percent) in natural samples.
 Category: Sulfur
| references

Sulfur or sulphur (/[invalid input: 'icon']ˈsʌlfər/ SUL-fər; see spelling below) is the chemical element with atomic number 16, represented by the symbol S. It is an abundant, multivalent non-metal. At normal conditions, sulfur atoms form cyclic octatomic molecules with chemical formula S8. Elemental sulfur is a bright yellow crystalline solid. Chemically, sulfur can react as either an oxidant or reducing agent. It oxidizes most metals and several nonmetals, including carbon, which leads to its negatives charge in most organosulfur compounds, but it reduces several strong oxidants, such as oxygen and fluorine.

In nature, sulfur can be found as the pure element and as sulfide and sulfate minerals. Elemental sulfur crystals are commonly sought after by mineral collectors for their brightly colored polyhedron shapes. Being abundant in native form, sulfur was known in ancient times, mentioned for its uses in ancient Greece, China and Egypt. Sulfur fumes were used as fumigants, and sulfur-containing medicinal mixtures were used as balms and antiparasitics. Sulfur is referenced in the Bible as brimstone, with this name still used in several nonscientific terms.[7] Sulfur was considered important enough to receive its own alchemical symbol. It was needed to make black gunpowder, and the bright yellow powder was hypothesized by alchemists to contain some of the properties of gold, which they sought to synthesize from it. In 1777, Antoine Lavoisier helped convince the scientific community that sulfur was a basic element, rather than a compound.

Elemental sulfur was once extracted from salt domes where it sometimes occurs in nearly pure form, but this method has been obsolete since the late 20th century. Today, almost all elemental sulfur is produced as a byproduct of removing sulfur-containing contaminants from natural gas and petroleum. The element's commercial uses are primarily in fertilizers, because of the relatively high requirement of plants for it, and in the manufacture of sulfuric acid, a primary industrial chemical. Other well-known uses for the element are in matches, insecticides and fungicides. Many sulfur compounds are odiferous, and the smell of odorized natural gas, skunk scent, grapefruit, and garlic is due to sulfur compounds. Hydrogen sulfide produced by living organisms imparts the characteristic odor to rotting eggs and other biological processes.

Sulfur is an essential element for all life, and is widely used in biochemical processes. In metabolic reactions, sulfur compounds serve as both fuels and respiratory (oxygen-replacing) materials for simple organisms. Sulfur in organic form is present in the vitamins biotin and thiamine, the latter being named for the Greek word for sulfur. Sulfur is an important part of many enzymes and also in antioxidant molecules like glutathione and thioredoxin. Organically bonded sulfur is a component of all proteins, as the amino acids cysteine and methionine. Disulfide bonds are largely responsible for the mechanical strength and insolubility of the protein keratin, found in outer skin, hair, and feathers, and the element contributes to their pungent odor when burned.

Characteristics

When burned, sulfur melts to a blood-red liquid and emits a blue flame which is best observed in the dark.

Physical

The first and the second ionization energies of sulfur are 999.6 and 2252 kJ·mol−1, respectively. Despite such high figures, oxidation of sulfur to +2 is rare, with more respect to +4 and +6 states. The fourth and sixth ionization energies are 4556 and 8495.8 kJ·mol−1, with height of the figures caused due to electron orbital transfer. These state are only stable with such strong oxidants as fluorine, oxygen and chlorine.

Sulfur forms polyatomic molecules with different chemical formulae, with the best-known allotrope being octasulfur, cyclo-S8. Octasulfur is a soft, bright-yellow solid with only a faint odor, similar to that of matches.[8] It melts at 115.21 °C, boils at 444.6 °C and sublimes easily.[7] At 95.2 °C, below its melting temperature, cyclo-octasulfur changes from α-octasulfur to β-polymorph.[9] The structure of the S8 ring is virtually unchanged by this phase change, which affects the intermolecular interactions. Between its melting and boiling temperatures, octasulfur changes its allotrope again, turning from β-octasulfur to γ-sulfur, again accompanied by a lowering in density but increasing in viscosity due to the formation of polymers.[9] At even higher temperatures, however, the viscosity decreases as depolymerization occurs. Molten sulfur assumes a dark red color above 200 °C.

Its density is roughly equal to 2 g·cm−3, depending on the allotrope. All of its stable allotropes are excellent electrical insulators.

Chemical

Sulfur burns with a blue flame concomitant with formation of sulfur dioxide, notable for its peculiar suffocating odor. Sulfur is insoluble in water, but soluble in carbon disulfide;— and to a lesser extent in other nonpolar organic solvents such as benzene and toluene.

Polymeric sulfur

Amorphous or "plastic" sulfur is produced by rapid cooling of molten sulfur—for example, by pouring it into cold water. X-ray crystallography studies show that the amorphous form may have a helical structure with eight atoms per turn. The long coiled polymeric molecules cause the brownish substance to be elastic, and in bulk this form has the feel of crude rubber. This form is metastable at room temperature and gradually reverts to crystalline molecular allotrope, which is no longer elastic. This process happens within a matter of hours to days, but can be rapidly catalyzed.

Allotropes

The structure of the cyclooctasulfur molecule, S8.

Sulfur forms more than 30 solid allotropes, more than any other element.[10] Besides S8, several other rings are known.[11] Removing one atom from the crown gives S7, which is more deeply yellow than S8. HPLC analysis of "elemental sulfur" reveals an equilibrium mixture of mainly S8, but also S7 and small amounts of S6.[12] Larger rings have been prepared, including S12 and S18.[13][14]

Isotopes

Sulfur has 25 known isotopes, four of which are stable: 32S (95.02%), 33S (0.75%), 34S (4.21%), and 36S (0.02%). Other than 35S, the radioactive isotopes of sulfur are all short lived. 35S is formed from cosmic ray spallation of 40Ar in the atmosphere. It has a half-life of 87 days.

When sulfide minerals are precipitated, isotopic equilibration among solids and liquid may cause small differences in the δS-34 values of co-genetic minerals. The differences between minerals can be used to estimate the temperature of equilibration. The δC-13 and δS-34 of coexisting carbonates and sulfides can be used to determine the pH and oxygen fugacity of the ore-bearing fluid during ore formation.

In most forest ecosystems, sulfate is derived mostly from the atmosphere; weathering of ore minerals and evaporites also contribute some sulfur. Sulfur with a distinctive isotopic composition has been used to identify pollution sources, and enriched sulfur has been added as a tracer in hydrologic studies. Differences in the natural abundances can also be used in systems where there is sufficient variation in the 34S of ecosystem components. Rocky Mountain lakes thought to be dominated by atmospheric sources of sulfate have been found to have different δ34S values from lakes believed to be dominated by watershed sources of sulfate.

Natural occurrence

Most of the yellow and orange hues of Io are due to elemental sulfur and sulfur compounds, produced by active volcanoes.

The most common isotope of sulfur, sulfur-32, is created in extremely large, extremely hot (over 2.5 billion kelvin) stars. This requires fusion of one nucleus of silicon plus one nucleus of helium.[15] For details of the process, see silicon burning. Partly because this so-called alpha process produces elements in abundance, sulfur is the 10th most common element in the universe. Sulfur, usually as sulfide, is present in many types of meteorites. Ordinary chondrites contain on average 2.1% sulfur, and carbonaceous chondrites may contain as much as 6.6%. Sulfur in meteorites is normally present as troilite (FeS), but other sulfides are found in some meteorites, and carbonaceous chondrites contain free sulfur, sulfates, and possibly other sulfur compounds.[16] The distinctive colors of Jupiter's volcanic moon, Io, are attributed to various forms of molten, solid and gaseous sulfur.[17]

Native sulfur crystals

On Earth, elemental sulfur can be found near hot springs and volcanic regions in many parts of the world, especially along the Pacific Ring of Fire. Such volcanic deposits are currently mined in Indonesia, Chile, and Japan. Sicily is also famous for its sulfur mines. Sulfur deposits are polycrystalline, and the largest documented single crystal measured 22×16×11 cm.[18][19]

A man carrying sulfur blocks from Kawah Ijen, a volcano in East Java, Indonesia (photo 2009)

Significant deposits of elemental sulfur also exist in salt domes along the coast of the Gulf of Mexico, and in evaporites in eastern Europe and western Asia. The sulfur in these deposits is believed to come from the action of anaerobic bacteria on sulfate minerals, especially gypsum, a process which is still seen today. However, native sulfur may be produced by geological processes alone, without the aid of living organisms (see below). Fossil-based sulfur deposits from salt domes have, until recently, been the basis for commercial production in the United States, Poland, Russia, Turkmenistan, and Ukraine.[20] As noted below, such sources are now of secondary commercial importance, and most are no longer worked.

Common naturally-occurring sulfur compounds include the sulfide minerals, such as pyrite (iron sulfide), cinnabar (mercury sulfide), galena (lead sulfide), sphalerite (zinc sulfide) and stibnite (antimony sulfide); and the sulfates, such as gypsum (calcium sulfate), alunite (potassium aluminium sulfate), and barite (barium sulfate). On Earth, just as upon Jupter's moon Io, elemental sulfur occurs naturally in volcanic emissions, including emissions from hydrothermal vents.

Production

Sulfur may be found as a pure mineral ("native" sulfur) and historically was usually obtained in this way. However, today's sulfur production is as a side product of other industrial processes, such as oil refining. In these products and processes, sulfur often occurs as undesired or detrimental sulfur compounds, which are extracted and converted to elemental sulfur.

As a mineral, native sulfur under salt domes is thought to be a fossil mineral resource, produced by the action of ancient bacteria on sulfate deposits. It was removed from such salt-dome mines mainly by the Frasch process.[20] In this method, superheated water was pumped into a native sulfur deposit to melt the sulfur, and then compressed air returned the relatively pure (99.5%) melted product to the surface. Throughout the 20th century this procedure produced elemental sulfur which required no further purification. However, due to a limited number of such sulfur deposits and the high cost of working them, this process for mining sulfur has not been employed in a major way anywhere in the world, since 2002.[21][22]

Sulfur recovered from hydrocarbons in Alberta, stockpiled for shipment in North Vancouver, B.C.

Today, sulfur is produced from petroleum, natural gas, and related fossil resources, from which it is obtained mainly as hydrogen sulfide (a gas). organosulfur compounds, which are undesirable impurities in petroleum, may be upgraded by subjecting them to hydrodesulfurization, which cleaves the C-S bonds:[21][22]

R-S-R + 2 H2 → 2 RH + H2S

The resulting hydrogen sulfide from this process, and also as it occurs in natural gas, is converted into elemental sulfur by the Claus process. This process entails oxidation of some hydrogen sulfide to sulfur dioxide and then the comproportionation of the two:[21][22]

3 O2 + 2 H2S → 2 SO2 + 2 H2O
SO2 + 2 H2S → 3 S + 2 H2O

Owing to the high sulfur content of the Athabasca Oil Sands, stockpiles of elemental sulfur from this process now exist throughout Alberta, Canada.[23]

Compounds

See also Category: sulfur compounds

Common oxidation states of sulfur range from −2 to +6. Sulfur forms stable compounds with all elements except the noble gases.

Sulfides

Treatment of sulfur with hydrogen gives hydrogen sulfide. When dissolved in water, hydrogen sulfide is mildly acidic:[7]

H2S HS + H+

Reduction of elemental sulfur gives polysulfides, which consist of chains of sulfur atoms terminated with S centers:

2 Na + S8 → Na2S8

This reaction highlights arguably the single most distinctive property of sulfur: its ability to catenate (bind to itself by formation of chains). Protonation of these polysulfide anions gives the polysulfanes, H2Sx where x = 2, 3, and 4.[24] Ultimately reduction of sulfur gives sulfide salts:

16 Na + S8 → 8 Na2S

The interconversion of these species is exploited in the sodium-sulfur battery. The radical anion S2 gives the blue color to the mineral lapis lazuli.

Lapis lazuli owes its blue color to a sulfur radical.

Elemental sulfur can also be oxidized, for example to give bicyclic S82+.

Oxides and oxyanions

The principal sulfur oxides are obtained by burning sulfur:

S + O2 → SO2
2 SO2 + O2 → 2 SO3

Other oxides are known, e.g. sulfur monoxide and disulfur mono- and dioxides, but they are unstable.

The sulfur oxides form numerous oxyanions with the formula SOn2–. Sulfur dioxide and sulfites (SO2−
3
) are related to the unstable sulfurous acid (H2SO3). Sulfur trioxide and sulfates (SO2−
4
) are related to sulfuric acid. Sulfuric acid and SO3 combine to give oleum, a solution of pyrosulfuric acid (H2S2O7) in sulfuric acid.

Peroxydisulfuric acid

Peroxides convert sulfur into unstable such as S8O, a sulfoxide. Peroxymonosulfuric acid (H2SO5) and peroxydisulfuric acids (H2S2O8), made from the action of SO3 on concentrated H2O2, and H2SO4 on concentrated H2O2 respectively.

The sulfate anion, SO2−
4

Thiosulfate salts (S
2
O2−
3
), sometimes referred as "hyposulfites", used in photographic fixing (HYPO) and as reducing agents, feature sulfur in two oxidation states. Sodium dithionite, (S
2
O2−
4
), contains the more highly reducing dithionite anion. Sodium dithionate (Na2S2O6) is the first member of the polythionic acids (H2SnO6), where n can range from 3 to many.

Halides and oxyhalides

Sulfur hexafluoride, SF6 is a dense gas that is used as nonreactive and nontoxic propellant. In contrast, sulfur tetrafluoride is a rarely used reagent that is highly toxic. The two main sulfur chlorides are sulfur dichloride (SCl2) and sulfur monochloride (S2Cl2). Sulfuryl chloride (SO2Cl2) and chlorosulfuric acid (ClSO3H) are derivatives of sulfuric acid. Thionyl chloride (SOCl2) is a common reagent in organic synthesis.

Pnictides

The most important S-N compound is the cage tetrasulfur tetranitride (S4N4). Heating this compound gives Polymeric sulfur nitride ((SN)x), which has metallic properties even though it does not contain any metal atoms. Thiocyanates contain the SCN group. Oxidation of thiocyanate gives thiocyanogen, (SCN)2 with the connectivity NCS-SCN. Phosphorus sulfides are numerous, the most important commercially being the cages P4S10 and P4S3.[25][26]

Metal sulfides

Many if not most minerals occur as sulfides. The principal ores of copper, zinc, nickel, cobalt, molybdenum and others are sulfides. These materials tend to be dark-colored semiconductors that are not readily attacked by water or even many acids. They are formed, both geochemically and in the laboratory, by the reaction of hydrogen sulfide with metal salts to form the metal sulfides. The mineral Galena (PbS) was the first demonstrated semiconductor and found a use as a signal rectifier in the cat's whiskers of early crystal radios. The iron sulfide called pyrite, the so-called "fool's gold," has the formula FeS2.[27] The upgrading of these ores, usually by roasting, is costly and environmentally hazardous. Sulfur corrodes many metals via the process called tarnishing.

Organic compounds

Some of the main classes of sulfur-containing organic compounds include the following (R, R', and R are organic groups such as CH3):[28]

  • Thioethers have the form R-S-R′. These compounds are the sulfur equivalents of ethers.
  • Sulfonium ions have the formula RR'R"S+, i.e. where three groups are attached to the cationic sulfur center. Dimethylsulfoniopropionate (DMSP; (CH3)2S+CH2CH2COO) is a sulfonium ion, which is important in the marine organic sulfur cycle.
  • Thiols, known also as "mercaptans" because of their mercury capturing property as chelators, have the form R-SH. These are the sulfur equivalents of alcohols. Treatment of thiols with base gives thiolates ions (R-S.
  • Sulfoxides and sulfones have the form R-S(=O)-R′ and R-S(=O)(=O)-R′. The simplest sulfoxide, DMSO, is a common solvent. A common sulfone is sulfolane C4H8SO2.
  • Sulfonic acids (R-SO3) are used in many detergents.

Some "inorganic" (in the sense of non-hydrogen-containing) carbon-sulfur compounds are known. Carbon disulfide (CS2), a volatile colorless liquid at standard conditions, is structurally similar to carbon dioxide. It is used as a solvent to make polymers. Whereas carbon monoxide is a highly stable gas, carbon monosulfide (CS) is a laboratory curiosity with only a fleeting existence.

Organosulfur compounds are responsible for the some of the unpleasant odors of decaying organic matter. Thiols and sulfides are used in the odoration of natural gas, notably, t-butyl mercaptan. The odor of garlic and "skunk stink" are also caused by organosulfur compounds. Not all organic sulfur compounds smell unpleasant at all concentrations; for example, grapefruit mercaptan, a sulfur-containing monoterpenoid is responsible for the characteristic scent of grapefruit. It should be noted that this thiol is present in very low concentrations. In larger concentrations, the odor of this compound is that typical of all thiols.

Finally, sulfur can be used in organics as a structural component to harden synthetic polymers, in a way similar to the biological use of disulfide bridges to reinforce proteins (see biological below). In the most common type of industrial "curing" or hardening and strengthening of natural rubber, elemental sulfur is heated with the rubber to the point that chemical reactions form disulfide bridges between isoprene units of the polymer. This process, patented in 1843, historically changed rubber into a major industrial product. The process was named vulcanization after the Roman god of the forge and volcanism, in honor of both the heat and sulfur used. Although vulcanization is applied to other polymers, and sometimes with crosslinking agents other than sulfur, variants of sulfur/rubber vulcanization continue to be used in producing automobile tires and other elastomer products.

History

Antiquity

Being abundantly available in native form, sulfur (Sanskrit, गन्धक sulvari; Latin sulphur, ) was known in ancient times and is referred to in the Torah (Genesis). English translations of the Bible commonly referred to burning sulfur as "brimstone", giving rise to the name of 'fire-and-brimstone' sermons, in which listeners are reminded of the fate of eternal damnation that await the unbelieving and unrepentant. It is from this part of the Bible that Hell is implied to "smell of sulfur" (likely due to its association with volcanic activity). According to the Ebers Papyrus, a sulfur ointment was used in ancient Egypt to treat granular eyelids. Sulfur was used for fumigation in preclassical Greece;[29] this is mentioned in the Odyssey.[30] Pliny the Elder discusses sulfur in book 35 of his Natural History, saying that its best-known source is the island of Melos. He also mentions its use for fumigation, medicine, and bleaching cloth.[31]

A natural form of sulfur known as shiliuhuang was known in China since the 6th century BC and found in Hanzhong.[32] By the 3rd century, the Chinese discovered that sulfur could be extracted from pyrite.[32] Chinese Daoists were interested in sulfur's flammability and its reactivity with certain metals, yet its earliest practical uses were found in traditional Chinese medicine.[32] A Song Dynasty military treatise of 1044 AD described different formulas for Chinese black powder, which is a mixture of potassium nitrate (KNO
3
), charcoal, and sulfur.

Early alchemists gave sulfur its own alchemical symbol which was a triangle at the top of a cross. In traditional medical skin treatment which predates modern era of scientific medicine, elemental sulfur has been used mainly as part of creams to alleviate various conditions such as scabies, ringworm, psoriasis, eczema and acne. The mechanism of action is not known, although elemental sulfur does oxidize slowly to sulfurous acid, which in turn (through the action of sulfite) acts as a mild reducing and antibacterial agent.[33][34][35]

Modern times

In 1777, Antoine Lavoisier helped convince the scientific community that sulfur was an element and not a compound. The Sicilian process was used in ancient times to obtain sulfur from rocks present in volcanic regions of Sicily. In this process, the sulfur deposits are piled and stacked in brick kilns built on sloping hillsides, and with airspaces between them. Then powdered sulfur is put on top of the sulfur deposit and ignited. As the sulfur burns, the heat melts the sulfur deposits, causing the molten sulfur to flow down the sloping hillside.

Sicilian kiln used to obtain sulfur from volcanic rock.

In 1867, sulfur was discovered in underground deposits in Louisiana and Texas. The highly successful Frasch process was developed to extract this resource.[36]

In the late 18th century, furniture makers used molten sulfur to produce decorative inlays in their craft. Because of the sulfur dioxide produced during the process of melting sulfur, the craft of sulfur inlays was soon abandoned. Molten sulfur is sometimes still used for setting steel bolts into drilled concrete holes where high shock resistance is desired for floor-mounted equipment attachment points. Pure powdered sulfur was also used as a medicinal tonic and laxative.[20]

Spelling and etymology

The element was traditionally spelt sulphur in the United Kingdom (since the 14th century),[37] most of the Commonwealth including India, Malaysia, South Africa, and Hong Kong, along with the rest of the Caribbean and Ireland. Sulfur is used in the United States, while both spellings are used in Canada and the Philippines.[37]

However, the IUPAC adopted the spelling sulfur in 1990, as did the Royal Society of Chemistry Nomenclature Committee in 1992.[38] The Qualifications and Curriculum Authority for England and Wales recommended its use in 2000,[39] and it now appears in GCSE exams.[40] The Oxford Dictionaries note that "In chemistry... the -f- spelling is now the standard form in all related words in the field in both British and US contexts"[41]

In Latin, the word is variously written sulpur, sulphur, and sulfur (the Oxford Latin Dictionary lists the spellings in this order). It is an original Latin name and not a Classical Greek loan, so the ph variant does not denote the Greek letter φ (phi). Sulfur in Greek is thion (θείον), whence comes the prefix thio-. The simplification of the Latin words p or ph to an f appears to have taken place towards the end of the classical period.[42][43]

Applications

Sulfuric acid

Elemental sulfur is mainly used as a precursor to other chemicals. Approximately 85% (1989) is converted to sulfuric acid (H2SO4):

2 S + 3 O2 + 2 H2O → 2 H2SO4

With sulfuric acid being central importance to the world's economies, its production and consumption is an indicator of a nation's industrial development.[44] For example with 36.1 million metric tons in 2007, the United States produces more sulfuric acid every year than any other inorganic industrial chemical.[45] The principal use for the acid is the extraction of phosphate ores for the production of fertilizer manufacturing. Other applications of sulfuric acid include oil refining, wastewater processing, and mineral extraction.[20]

Sulfuric acid production in 2000

Other large scale sulfur chemicals

Sulfur reacts directly with methane to give carbon disulfide, which is used to manufacture cellophane and rayon.[20] One of the direct uses of sulfur is in vulcanization of rubber, where polysulfides crosslink organic polymers. Sulfites are heavily used to bleach paper. Sulfites are also used as preservatives in dried fruit. Many surfactants and detergents, e.g. sodium lauryl sulfate, are produced are sulfate derivatives. Calcium sulfate, gypsum, (CaSO4·2H2O) is mined on the scale of 100 million tons each year for use in Portland cement and fertilizers.

When silver-based photography was widespread, sodium and ammonium thiosulfate were widely used as "fixing agents." Sulfur is a component of gunpowder.

Fertilizer

Sulfur is increasingly used as a component of fertilizers. The most important form of sulfur for fertilizer is the mineral calcium sulfate. Elemental sulfur is hydrophobic (that is, it is not soluble in water) and therefore cannot be directly utilized by plants. Over time, soil bacteria can convert it to soluble derivatives which can then be utilized by plants. Sulfur also improves the use efficiency of other essential plant nutrients, particularly nitrogen and phosphorus.[46] Biologically produced sulfur particles are naturally hydrophilic due to a biopolymer coating. This sulfur is therefore easier to disperse over the land (via spraying as a diluted slurry), and results in a faster release.

Plant requirements for sulfur are equal to or exceed those for phosphorus. It is one of the major nutrients essential for plant growth, root nodule formation of legumes and plants protection mechanisms. Sulfur deficiency has become widespread in many countries in Europe.[47][48][49] Because atmospheric inputs of sulfur will continue to decrease, the deficit in the sulfur input/output is likely to increase, unless sulfur fertilizers are used.

Fine chemicals

A molecular model of the pesticide malathion.

Organosulfur compounds are also used in pharmaceuticals, dyestuffs, and agrochemicals. Many drugs contain sulfur, early examples being the sulfa drugs. Sulfur is a part of many bacterial defense molecules. Most beta lactam antibiotics, including the penicillins, cephalosporins and monolactams contain sulfur.[28]

Magnesium sulfate, better known as Epsom salts, can be used as a laxative, a bath additive, an exfoliant, magnesium supplement for plants, or a desiccant.

Fungicide and pesticide

Elemental sulfur is one of the oldest fungicides and pesticides. Dusting sulfur, elemental sulfur in powdered form, is a common fungicide for grapes, strawberry, many vegetables and several other crops. It has a good efficacy against a wide range of powdery mildew diseases as well as black spot. In organic production, sulfur is the most important fungicide. It is the only fungicide used in organically farmed apple production against the main disease apple scab under colder conditions. Biosulfur (biologically produced elemental sulfur with hydrophilic characteristics) can be used well for these applications.

Standard-formulation dusting sulfur is applied to crops with a sulfur duster or from a dusting plane. Wettable sulfur is the commercial name for dusting sulfur formulated with additional ingredients to make it water miscible.[50][51] It has similar applications and is used as a fungicide against mildew and other mold-related problems with plants and soil.

Sulfur is also used as an "organic" (i.e. "green") insecticide (actually an acaricide) against ticks and mites. A common method of use is to dust clothing or limbs with sulfur powder. Some livestock owners set out a sulfur salt block as a salt lick. [citation needed]

Biological role

Protein and organic cofactors

Sulfur is an essential component of all living cells. It is the seventh or eighth most abundant element in the human body by weight, being about as common as potassium, and a little more common than sodium or chlorine. A 70 kg human body contains about 140 grams of sulfur.

In plants and animals, the amino acids cysteine and methionine contain most of the sulfur. The element is thus present in all polypeptides, proteins, and enzymes that contain these amino acids. Disulfide bonds (S-S bonds) formed between cysteine residues in peptide chains are very important in protein assembly and structure. These covalent bonds between peptide chains confer extra toughness and rigidity.[52] For example, the high strength of feathers and hair is in part due to their high content of S-S bonds and their high content of cysteine and sulfur. Eggs are high in sulfur because large amounts of the element are necessary for feather formation, and the characteristic odor of rotting eggs is due to hydrogen sulfide. The high disulfide bond content of hair and feathers contributes to their indigestibility, and also to their characteristic disagreeable odor when burned.

Homocysteine and taurine are other sulfur-containing acids that are similar in structure, but which are not coded by DNA, and are not part of the primary structure of proteins. Many important cellular enzymes use prosthetic groups ending with -SH moieties to handle reactions involving acyl-containing biochemicals: two common examples from basic metabolism are coenzyme A and alpha-lipoic acid.[52] Two of the 13 classical vitamins, biotin and thiamine contain sulfur, with the latter being named for its sulfur content. Sulfur plays an important part, as a carrier of reducing hydrogen and its electrons, for cellular repair of oxidation. Reduced glutathione, a sulfur-containing tripeptide, is a reducing agent through its sulfhydryl (-SH) moiety derived from cysteine. The thioredoxins are essential classes of small proteins acting as general reducing agents in cells, that use pairs of reduced cysteines to similar effect.

Methanogenesis, the route to most of the world's methane, is a multistep biochemical transformation of carbon dioxide. This conversion requires several organosulfur cofactors. These include coenzyme M, CH3SCH2CH2SO3, the immediate precursor to methane.[53]

Metalloproteins and inorganic cofactors

Inorganic sulfur forms a part of iron-sulfur clusters as well as many copper, nickel, and iron proteins. Most pervasive are the ferrodoxins, which serve as electron shuttles in cells. In bacteria, the important nitrogenase enzymes contains an Fe-Mo-S cluster, is a catalyst that performs the important function of nitrogen fixation, converting atmospheric nitrogen to ammonia that can be used by microorganisms and plants to make proteins, DNA, RNA, alkaloids, and the other organic nitrogen compounds necessary for life.[54]

Sulfur metabolism

Sulfur may also serve as energy (chemical food) source for bacteria that use hydrogen sulfide (H2S) in the place of water as the electron donor in a primitive photosynthesis-like process in which oxygen is the electron receptor. The photosynthetic green and purple sulfur bacteria and some chemolithotrophs use elemental oxygen to carry out such oxidization of hydrogen sulfide to produce elemental sulfur (So), oxidation state = 0. Primitive bacteria which live around deep ocean volcanic vents oxidize hydrogen sulfide in this way with oxygen: see giant tube worm for an example of large organisms (via bacteria) making metabolic use of hydrogen sulfide as food to be oxidized.

The so-called sulfur bacteria, by contrast, "breathe sulfate" instead of oxygen. They use sulfur as the electron acceptor, and reduce various oxidized sulfur compounds back into sulfide, often into hydrogen sulfide. They also can grow on a number of other partially oxidized sulfur compounds (e. g. thiosulfates, thionates, polysulfides, sulfites). The hydrogen sulfide produced by these bacteria is responsible for some of the smell of intestinal gases (flatus) and decomposition products.

Sulfur is absorbed by plants via the roots from soil as the sulfate and transported as a phosphate ester. Sulfate is reduced to sulfide via sulfite before it is incorporated into cysteine and other organosulfur compounds.[55]

SO42– → SO32– → H2S → cysteine

Precautions

Elemental sulfur is non-toxic, but it can burn, producing sulfur dioxide. Although sulfur dioxide is sufficiently safe to be used as a food additive in small amounts, at high concentrations it harms the lungs, eyes or other tissues. In organisms without lungs such as insects or plants, it otherwise prevents respiration. Sulfur trioxide and sulfuric acid are similarly highly corrosive, due to the strong acids that form on contact with water.

Effect of acid rain on a forest, Jizera Mountains, Czech Republic

The burning of coal and/or petroleum by industry and power plants generates sulfur dioxide (SO2), which reacts with atmospheric water and oxygen to produce sulfuric acid (H2SO4) and sulfurous acid (H2SO3). These acids are components of acid rain, which lower the pH of soil and freshwater bodies, sometimes resulting in substantial damage to the environment and chemical weathering of statues and structures. Fuel standards increasingly require sulfur to be extracted from fossil fuels to prevent the formation of acid rain. This extracted sulfur is then refined and represents a large portion of sulfur production. In coal fired power plants, the flue gases are sometimes purified. In more modern power plants that use syngas the sulfur is extracted before the gas is burned.

Hydrogen sulfide is as toxic as hydrogen cyanide and kills by the same mechanism, although hydrogen sulfide is less likely to result in surprise poisonings from small inhaled amounts, due to its more disagreeable warning odor. However, although very pungent at first awareness to the human nose, hydrogen sulfide quickly deadens the sense of smell, so potential victims breathing larger and larger quantities of it may be unaware of its presence until severe symptoms occur (these can then quickly lead to death).

See also

References

  1. ^ "Standard Atomic Weights: Sulfur". CIAAW. 2009.
  2. ^ 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.
  3. ^ a b Arblaster, John W. (2018). Selected Values of the Crystallographic Properties of Elements. Materials Park, Ohio: ASM International. ISBN 978-1-62708-155-9.
  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. ^ Weast, Robert (1984). CRC, Handbook of Chemistry and Physics. Boca Raton, Florida: Chemical Rubber Company Publishing. pp. E110. ISBN 0-8493-0464-4.
  6. ^ "Sulfur History". Georgiagulfsulfur.com. Retrieved 2022-02-12.
  7. ^ a b c Greenwood, N. N.; & Earnshaw, A. (1997). Chemistry of the Elements (2nd Edn.), Oxford:Butterworth-Heinemann. ISBN 0-7506-3365-4.
  8. ^ A strong odor called "smell of sulfur" actually is given off by several sulfur compounds, like hydrogen sulfide and organosulfur compounds.
  9. ^ a b Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. pp. 645–662. ISBN 978-0-08-037941-8.
  10. ^ Steudel, Ralf; Eckert, Bodo (2003). "Solid Sulfur Allotropes Sulfur Allotropes". Topics in Current Chemistry. 230: 1–80. doi:10.1007/b12110.
  11. ^ Steudel, R. (1982). "Homocyclic Sulfur Molecules". Topics in Current Chemistry. 102: 149–176. doi:10.1007/3-540-11345-2_10.
  12. ^ Tebbe, Fred N.; Wasserman, E.; Peet, William G.; Vatvars, Arturs; Hayman, Alan C. (1982). "Composition of Elemental Sulfur in Solution: Equilibrium of S
    6
    , S7, and S8 at Ambient Temperatures". Journal of the American Chemical Society. 104 (18): 4971. doi:10.1021/ja00382a050.
  13. ^ Meyer, Beat (1964). "Solid Allotropes of Sulfur". Chemical Reviews. 64 (4): 429–451. doi:10.1021/cr60230a004.
  14. ^ Meyer, Beat (1976). "Elemental sulfur". Chemical Reviews. 76 (3): 367–388. doi:10.1021/cr60301a003.}
  15. ^ Cameron, A. G. W. (1957). "Stellar Evolution, Nuclear Astrophysics, and Nucleogenesis" (PDF). CRL-41.
  16. ^ Mason, B. (1962). Meteorites. New York: John Wiley & Sons. p. 160. ISBN 0908678843.
  17. ^ Lopes, Rosaly M C; Williams, David A (2005). "Io after Galileo". Reports on Progress in Physics. 68 (2): 303–340. doi:10.1088/0034-4885/68/2/R02.
  18. ^ Rickwood, P. C. (1981). "The largest crystals" (PDF). American Mineralogist. 66: 885–907.
  19. ^ "The giant crystal project site". Retrieved 2009-06-06.
  20. ^ a b c d e Nehb, Wolfgang (2006). "Sulfur". Ullmann's Encyclopedia of Industrial Chemistry. Wiley-VCH Verlag. doi:10.1002/14356007.a25_507.pub2. {{cite encyclopedia}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  21. ^ a b c Eow, John S. (2002). "Recovery of sulfur from sour acid gas: A review of the technology". Environmental Progress. 21 (3): 143. doi:10.1002/ep.670210312.
  22. ^ a b c Schreiner, Bernhard (2008). "Der Claus-Prozess. Reich an Jahren und bedeutender denn je". Chemie in unserer Zeit. 42 (6): 378. doi:10.1002/ciuz.200800461.
  23. ^ Hyndman, A. W.; Liu, J. K.; Denney, D. W. (1982). "Sulfur Recovery from Oil Sands". 183: 69. doi:10.1021/bk-1982-0183.ch005. {{cite journal}}: Cite journal requires |journal= (help)
  24. ^ Handbook of Preparative Inorganic Chemistry, 2nd Ed. Edited by G. Brauer, Academic Press, 1963, NY. Vol. 1. p. 421.
  25. ^ Heal, H. G. (1980). The Inorganic Heterocyclic Chemistry of Sulfur, Nitrogen, and Phosphorus. London: Academic Press. ISBN 0123356806.
  26. ^ Chivers, T. (2004). A Guide To Chalcogen-Nitrogen Chemistry. Singapore: World Scientific. ISBN 9812560955.
  27. ^ Vaughan, D. J.; Craig, J. R. "Mineral Chemistry of Metal Sulfides" Cambridge University Press, Cambridge (1978) ISBN 0-521-21489-0
  28. ^ a b R. J. Cremlyn “An Introduction to Organosulfur Chemistry” John Wiley and Sons: Chichester (1996). ISBN 0 471 95512 4.
  29. ^ George Rapp Archaeomineralogy, 2nd ed., Springer: 2009, ISBN 978-3-540-78593-4 p. 242.
  30. ^ Odyssey, book 22, lines 480–495.
  31. ^ Pliny the Elder on science and technology, John F. Healy, Oxford University Press, 1999, ISBN 0198146876 pp. 247–249.
  32. ^ a b c Zhang Yunming (1986). "The History of Science Society: Ancient Chinese Sulfur Manufacturing Processes". Isis. 77 (3): 487. doi:10.1086/354207.
  33. ^ Lin, AN; Reimer, RJ; Carter, DM (1988). "Sulfur revisited". Journal of the American Academy of Dermatology. 18 (3): 553–8. doi:10.1016/S0190-9622(88)70079-1. PMID 2450900.
  34. ^ Maibach, HI; Surber, C; Orkin, M (1990). "Sulfur revisited". Journal of the American Academy of Dermatology. 23 (1): 154–6. doi:10.1016/S0190-9622(08)81225-X. PMID 2365870.
  35. ^ Gupta, AK; Nicol, K (2004). "The use of sulfur in dermatology". Journal of drugs in dermatology : JDD. 3 (4): 427–31. PMID 15303787.
  36. ^ Botsch, Walter (2001). "Chemiker, Techniker, Unternehmer: Zum 150. Geburtstag von Hermann Frasch". Chemie in unserer Zeit (in German). 35 (5): 324–331. doi:10.1002/1521-3781(200110)35:5<324::AID-CIUZ324>3.0.CO;2–9. {{cite journal}}: Check |doi= value (help)
  37. ^ a b Michie, C. A.; Langslow, D. R. (1988). "Sulphur or sulfur? A tale of two spellings". Britisch Medical Journal. 297 (6664): 1697. doi:10.1136/bmj.297.6664.1697.
  38. ^ McNaught, Alan (1991). "Journal style update". The Analyst. 116 (11): 1094. doi:10.1039/AN9911601094.
  39. ^ Sulphur, Worldwidewords
  40. ^ "General Certificate of Secondary Education (Science A Chemistry" (PDF). Foundation Tier and Higher Tier. March 2010. Retrieved 2011-02-27.
  41. ^ Ask Oxford, accessed 12 November 2010.
  42. ^ "Sulphuricum Sulphur". Vanderkrogt.net. Retrieved 2011-02-27.
  43. ^ Kelly, Donovan P. (1995). "Sulfur and its Doppelgänger". Arch. Microbiol. 163 (3): 157–158. doi:10.1007/BF00305347.
  44. ^ Sulfuric Acid Growth
  45. ^ Ober, Joyce A. "Mineral Yearbook 2007: Sulfur" (PDF). United States Geological Survey.
  46. ^ Sulfur as a fertilizer
  47. ^ Zhao, F (1999). "Sulphur Assimilation and Effects on Yield and Quality of Wheat". Journal of Cereal Science. 30 (1): 1. doi:10.1006/jcrs.1998.0241.
  48. ^ Blake-Kalff, M.M.A. (2000). Plant and Soil. 225 (1/2): 95. doi:10.1023/A:1026503812267. {{cite journal}}: Missing or empty |title= (help)
  49. ^ Ceccotti, S. P. (1996). "Plant nutrient sulphur-a review of nutrient balance, environmental impact and fertilizers". Fertilizer Research. 43 (1–3): 117. doi:10.1007/BF00747690.
  50. ^ Mohamed, Abdel-Mohsen Onsy; El Gamal, M. M (2010-07-13). Sulfur Concrete for the Construction Industry: A Sustainable Development Approach. pp. 104–105. ISBN 9781604270051.
  51. ^ Every, Richard L.; et al. (1968-08-20). "Method for Preparation of Wettable Sulfur" (PDF). Retrieved 2010-05-20. {{cite web}}: Explicit use of et al. in: |author= (help)
  52. ^ a b Nelson, D. L.; Cox, M. M. (2000.). Lehninger, Principles of Biochemistry (3rd ed.). New York: Worth Publishing. ISBN 1-57259-153-6. {{cite book}}: Check date values in: |year= (help)
  53. ^ Thauer, R. K. (1998). "Biochemistry of methanogenesis: a tribute to Marjory Stephenson:1998 Marjory Stephenson Prize Lecture". Microbiology. 144 (9): 2377. doi:10.1099/00221287-144-9-2377. PMID 9782487.
  54. ^ Lippard, S. J.; Berg, J. M. (1994). Principles of Bioinorganic Chemistry. University Science Books. ISBN 0-935702-73-3.
  55. ^ Heldt, Hans-Walter (1996). Pflanzenbiochemie. Heidelberg: Spektrum Akademischer Verlag. pp. 321–333. ISBN 3-8274-0103-8.

External links

Template:Link FA