Fluorine

oxygenfluorineneon
-

F

Cl
Appearance


Liquid fluorine at cryogenic temperatures
General properties
Name, symbol, number fluorine, F, 9
Pronunciation /ˈflʊərn/, /ˈflʊərɪn/, /ˈflɔːrn/
Element category halogen
Group, period, block 172, p
Standard atomic weight 18.9984032(5)[1]
Electron configuration 1s2 2s2 2p5[2]
Electrons per shell 2, 7[2] (Image)
Physical properties
Phase gas
Density (0 °C, 101.325 kPa)
1.696[3] g/L
Liquid density at b.p. 1.505[4] g·cm−3
Melting point 53.53 K, −219.62 °C, −363.32[5] °F
Boiling point 85.03 K, −188.12 °C, −306.62[5] °F
Critical point 144.00 K, 5.220[6] MPa
Heat of fusion 0.51[7] kJ·mol−1
Heat of vaporization 3.27[7] kJ·mol−1
Molar heat capacity (Cp) (21.1 °C) 825[8] J·mol−1·K−1
(Cv) (21.1 °C) 610[8] J·mol−1·K−1
Vapor pressure
P (Pa) 1 10 100 1 k 10 k 100 k
at T (K) 38 44 50 58 69 85
Atomic properties
Oxidation states −1
(oxidizes oxygen)
Electronegativity 3.98[9] (Pauling scale)
Ionization energies
(more)
1st: 1681.0[10] kJ·mol−1
2nd: 3374.2[10] kJ·mol−1
3rd: 6050.4[10] kJ·mol−1
Covalent radius 60[11] pm
Van der Waals radius 147[12] pm
Miscellanea
Crystal structure cubic
Crystal structure note the structure refers to solid fluorine, at boiling point, 1 atm[13]
Magnetic ordering diamagnetic[14]
Thermal conductivity 0.02591[15] W·m−1·K−1
CAS registry number 7782-41-4[2]
Most stable isotopes
Main article: Isotopes of fluorine
iso NA half-life DM DE (MeV) DP
18F trace 109.77 min β+ (97%) 0.633 18O
ε (3%) 1.656 18O
19F 100% 19F is stable with 10 neutrons
reference[16]
· r

Fluorine is the chemical element with atomic number 9, represented by the symbol F. It is the lightest element of the halogen column of the periodic table and has a single stable isotope, fluorine-19. At standard pressure and temperature, fluorine is a pale yellow gas composed of diatomic molecules, F2. In stars, fluorine is rare compared to other light elements. On Earth, fluorine is more common; it is the 13th most abundant element in the crust.

Fluorine's most important mineral, fluorite, was first formally described in 1530, in the context of metal smelting. The mineral's name derives from the Latin verb fluo, which means "stream" or "flow," because fluorite was added to metal ores to lower their melting points. Suggested to be a chemical element in 1811, "fluorine" was named after the source mineral. Several chemists (called the "fluorine martyrs") died in accidents while trying to isolate the element. In 1886, French chemist Henri Moissan succeeded in generating fluorine. His method of electrolysis remains the only industrial production method of fluorine gas. The main use of elemental fluorine, uranium enrichment, was developed during the Manhattan Project.

Because of the difficulty in making elemental fluorine, the vast majority of commercial fluorine is never reduced to the element. Fluorine has the highest electron affinity of any element but chlorine and is an extremely strong oxidizing agent. Fluorine forms stable compounds, fluorides, with all elements except helium and neon. Hydrofluoric acid, in contrast to hydrochloric acid, is only a weak acid, but it is nonetheless extremely corrosive. Fluorides of lighter metal elements are ionic compounds (salts). Heavier metal elements such as uranium form volatile coordination compounds (molecules with several fluorine atoms surrounding a metal atom).

Organic fluorine compounds tend to have high chemical and thermal stability and water-repellent properties. Several have large-scale commercial application, such as the fluorinated plastic polytetrafluoroethylene ("Teflon") used in cookware. Another major application is fluorinated refrigerants. Here, traditional chlorofluorocarbons ("Freons"), which cause ozone depletion, have been largely replaced by hydrofluorocarbons. A few organofluorine compounds are synthesized in microorganisms and plants. But, although it helps prevent tooth decay, fluorine is not an essential mineral for mammals. Several fluorine compounds, as well as elemental fluorine itself, are dangerously toxic. Nevertheless, an increasing number of pharmaceuticals (about 10% of new drugs) contain fluorine.

Characteristics

Electronic structure

A fluorine atom has nine protons and nine electrons, arranged in electronic configuration [He]2s22p5, one fewer than neon.[17] Fluorine's outer electrons are relatively separate from each other, and thus they do not shield each other from the nucleus. Therefore, they experience a relatively high effective nuclear charge. Because of this, fluorine is reluctant to ionize and has an attraction for one more electron to achieve the extremely stable neon-like arrangement.[17]

Fluorine's first ionization energy (energy required to remove an electron to form F+) is 1,681 kilojoules per mole, which is higher than for any other element except neon and helium.[10] The second and third ionization energies of fluorine are 3,374 and 6,147 kilojoules per mole, respectively.[10] Fluorine's electron affinity (energy released by adding an electron to form F) is 328 kilojoules per mole, which is higher than that of any other element except chlorine.[18] Fluorine has a relatively small covalent radius, about 60 picometers, which is more than neon and less than oxygen.[19]

Isotopes

Fluorine occurs naturally on Earth exclusively in the form of its only stable isotope, fluorine-19,[20] which makes the element monoisotopic and mononuclidic. Seventeen radioisotopes have been synthesized: mass numbers 14–18 and 20–31.[21] Fluorine-18 is the most stable radioisotope of fluorine, with a half-life of 109.77 minutes. It is also the lightest unstable nuclide with equal odd numbers of protons and neutrons.[22]

The lightest fluorine isotopes, 14–16, decay by electron capture. F-17 and F-18 undergo beta plus decay (positron emission). All isotopes heavier than the stable fluorine-19 decay by beta minus mode (electron emission). Some of them also decay by neutron emission.[21]

Only one nuclear isomer (long-lived excited nuclear state), fluorine-18m, has been characterized.[23] Its half-life before gamma ray emission is 160 nanoseconds. This is less than the decay half-life of any of the fluorine radioisotope nuclear ground states except numbers 14–16, 28, and 31.[23]

Physical properties

Fluorine form diatomic molecules that are gaseous at room temperature. The density is about 1.3 times that of air.[24][note 1] Though sometimes cited as yellow-green, fluorine gas is actually a very pale yellow. Its color can only be observed in concentrated fluorine gas when looking down the axis of long tubes. It appears transparent when observed from the side in normal glass tubes or if allowed to escape into the atmosphere.[26] The element has a "pungent" characteristic odor that is noticeable in concentrations as low as 20 ppb.[27]

Fluorine condenses to a bright yellow liquid at −188.1 °C (−306.6 °F),[5] a comparable temperature to the boiling points of oxygen and nitrogen. Fluorine solidifies at −219.6 °C (−363.3 °F)[5] into a cubic structure, called beta-fluorine. This phase is transparent and soft, with significant disorder of the molecules. At −227.5 °C (−377.5 °F) fluorine undergoes a solid–solid phase transition into a monoclinic structure called alpha-fluorine. This phase is opaque and hard with close-packed layers of molecules. The solid state phase change requires more energy than the melting point transition and can be violent, shattering samples and blowing out sample holder windows. In general, fluorine's solid state is more similar to oxygen than to the other halogens.[28][29]

Elemental fluorine
An 1892 observation of fluorine gas (middle), compared to air (left) and chlorine (right): the observation was done looking down 5 meter long tubes. 
Solid alpha-fluorine's crystal structure: fluorine molecules lie in shingled layers. 


Chemical reactivity

Halogen bond energies (kJ/mol)[30]
X XX HX BX3 AlX3 CX4
F 159 574 645 582 456
Cl 243 428 444 427 327
Br 193 363 368 360 272
I 151 294 272 285 239

Fluorine's chemistry is dominated by its tendency to gain an electron. It is the most electronegative element and a strong oxidant.[9] The removal of an electron from a fluorine atom requires so much energy that no known oxidant can oxidize fluorine to any positive oxidation state.[31]

Fluorine gas is highly reactive with other substances because of its oxidizing power—which leads to strong bonds with other atoms—and because of the relative weakness of the fluorine–fluorine bond. That bond energy is similar to the easily cleaved oxygen–oxygen bonds of peroxides or nitrogen–nitrogen bonds of hydrazines and significantly weaker than those of dichlorine or dibromine molecules.[32] The covalent radius of fluorine in difluorine molecules, about 71 picometers, is significantly larger than that in other compounds because of the weak bonding between fluorine atoms.[33]

External videos
Fluorine video from the University of Nottingham: Cold gas impinging on several substances causes bright flames.

Reactions with fluorine are often sudden or explosive.[24] Many generally non-reactive substances such as bricks, water, and frozen meat will burn with a bright flame in a jet of fluorine gas.[24]

Fluorine forms compounds, fluorides, with all elements except neon and helium. All of the elements up to einsteinium, element 99, have been checked except for astatine and francium.[34] Fluorine is also known to form compounds with rutherfordium, element 104,[35] and seaborgium, element 106.[36] Several heavy radioactive elements have not been fluoridated because of their extreme rarity, but such reactions are theoretically possible.[37]

All metals react with fluorine, but conditions vary with the metal. Often, the metal must be powdered because many metals passivate (form protective layers of the metal fluoride that resist further fluoridation). Alkali metals react with fluorine violently. Alkaline earth metals react at room temperature as well but do not release as much heat. The noble metals ruthenium, rhodium, palladium, platinum, and gold react least readily, requiring pure fluorine gas at 300–450 °C (575–850 °F).[38]

Fluorine reacts explosively with hydrogen in a manner similar to that of alkali metals.[39] The halogens react readily with fluorine gas[40] as does the heavy noble gas radon.[41] The lighter noble gases xenon and krypton can be made to react with fluorine under special conditions and argon will combine with hydrogen fluoride.[42] Nitrogen, with its very stable triple bonds, requires electric discharge and high temperatures to combine directly with fluorine.[43]

Origin and occurrence

In the universe

Abundance in the Solar System[44]
Atomic
number
Element Relative
amount
6 Carbon 4,800
7 Nitrogen 1,500
8 Oxygen 8,800
9 Fluorine 1
10 Neon 1,400
11 Sodium 24
12 Magnesium 430

From the perspective of cosmology, fluorine is relatively rare with 400 ppb in the universe. Within stars, any fluorine that is created is rapidly eliminated through nuclear fusion: either with hydrogen to form oxygen and helium, or with helium to make neon and hydrogen. The presence of fluorine at all—outside of temporary existence in stars—is somewhat of a mystery because of the need to escape these fluorine-destroying reactions.[45][46]

Three theoretical solutions to the mystery exist. In type II supernovae, atoms of neon are hit by neutrinos during the explosion and converted to fluorine. In Wolf-Rayet stars (blue stars over 40 times heavier than the Sun), a strong solar wind blows the fluorine out of the star before hydrogen or helium can destroy it. In asymptotic giant branch (a type of red giant) stars, fusion reactions occur in pulses and convection lifts fluorine out of the inner star. Only the red giant hypothesis has supporting evidence from observations.[45][46]

Even though fluorine, due to its chemical activity, does not exist in its elementary state on Earth, it can be found in the interstellar medium.[47] Fluorine cations have been seen in planetary nebulae and in stars, including our Sun.[48]

On Earth

Fluorine is the thirteenth most common element in Earth's crust, comprising between 600 and 700 ppm of the crust by mass. Due to its reactivity, it is found as fluoride ion rather than as the element. Three minerals exist that are industrial sources of fluorine.[49][50]

Notable fluorine-containing minerals
Fluorite
 
Fluorapatite
 
Cryolite
 

Several other minerals, such as the gemstone topaz, contain fluoride. Fluoride is not significant in seawater or brines, unlike the other halides, because the alkaline earth fluorides precipitate out of water.[50]

Organofluorines have been observed in volcanic eruptions and in geothermal springs. Their ultimate origin (physical formation under geological conditions or initial biological production and deposition in sediments) is unclear. They are not a commercially important source of fluorine, but are trace environmental contaminants whose amount is being studied.[52]

History

"Fluorine" is a word that ultimately derives from the Latin verb fluo, meaning "flow." The mineral fluorite, a natural form of calcium fluoride, was first mentioned in 1529 by Georgius Agricola, who named it after its use as a "flux"—an additive that helps melt ores and slags during smelting.[53][54] Agricola first named the mineral "fluorspar" as a latinization of the German Flußspat.[55] Since then, the mineral has been renamed "fluorite," although "fluorspar" is still sometimes used.[56]

In 1670, Heinrich Schwanhard, a German glass cutter, found that a vapor was produced when he treated "fluorspar" with strong acid. He used this property to develop a new art form, covering portions of glassware with protective varnish then exposing it to vapor, ending with clear figures on a cloudy background.[57][58]

Andreas Sigismund Marggraf made the first recorded preparation of "fluoric acid" (hydrofluoric acid in modern nomenclature) in 1764 when he heated fluorite with sulfuric acid in glass, which was greatly corroded by the product.[59] In 1771, Swedish chemist Carl Wilhelm Scheele repeated this reaction.[59] In 1810, French physicist André-Marie Ampère suggested that the acid was a compound of hydrogen with an unknown element, analogous to chlorine.[60] Fluorite was then shown to be mostly composed of calcium fluoride.[61][62]

Sir Humphry Davy originally suggested the name fluorine, taking the root from the name of "fluoric acid" and the -ine suffix, similarly to other halogens. This name, with modifications, came to most European languages. (However, Greek, Russian, and several other languages use the ftor or deratives, which comes from Greek φθόριος (phthorios), meaning "destructive.") The new Latin name (fluorum) gave the element its current symbol, F, although the symbol Fl is seen in early papers.[63]

Owing to its extreme reactivity, elemental fluorine was not isolated until many years after the characterization of fluorite. Progress in isolating elemental fluorine was slow because it could only be prepared electrolytically and because the gas reacted with most materials. The generation of elemental fluorine from fluorspar proved to be exceptionally dangerous, killing or blinding several early experimenters. Jean Dussaud referred to these men as "fluorine martyrs," a term still used.[62] Edmond Frémy decided that it might be preferable to pass the electric current through hydrofluoric acid instead. Unfortunately, hydrogen fluoride was only available in an aqueous solution. Frémy therefore devised a method for producing anhydrous hydrogen fluoride by acidifying potassium hydrogen fluoride (now known as potassium bifluoride or KHF2). Unfortunately, KHF2 will not pass an electric current.[57]

French chemist Henri Moissan, formerly one of Frémy's students, then took up the battle. After trying many different elements, he took Frémy's potassium bifluoride (created from hydrogen fluoride) and mixed it into a solution of hydrogen fluoride (so that the insulating potassium bifluoride could have an electric charge passed through the fluorides which formed the liquid of the solution) in equipment crafted from a mixture of platinum and iridium (even more non-reactive with fluorine than platinum), with flurospar stoppers (which wouldn't carry an electrical current and wouldn't react with fluorine). After 74 years of effort by other chemists, on 26 June 1886, Moissan reported the isolation of elemental fluorine.[57][64] He later discovered a less expensive method of producing fluorine, by using copper equipment overlain with copper fluoride. He received the 1906 Nobel Prize in chemistry for his fluorine isolation and invention of the electric arc furnace.[65]

"One can indeed make various hypotheses on the nature of the liberated gas; the simplest would be that we are in the presence of fluorine, but it would be possible, of course, that it might be a perfluoride of hydrogen or even a mixture of hydrofluoric acid and ozone..."

-Henri Moissan, to the French Academy (translated, italics in original)[66]

During the 1930s and 1940s, the DuPont company commercialized organofluorine compounds at large scales. Following trials of chlorofluorcarbons as refrigerants by researchers at General Motors, DuPont developed large-scale production of Freon-12. DuPont and GM started a joint venture in 1930 to market the new product, but soon DuPont took over the business. The substance proved to be a marketplace hit, rapidly replacing earlier, more toxic, refrigerants and growing the overall market for kitchen refrigerators.[59][67][68]

In 1938, polytetrafluoroethylene (DuPont brand name Teflon) was discovered by accident by a recently-hired DuPont Ph.D., Roy J. Plunkett. While working with tetrafluoroethylene gas, he noticed missing weight. Scraping down his container, he found white flakes of a new-to-the-world polymer. Tests showed the substance was resistant to corrosion from most substances and had better high temperature stability than any other plastic. By early 1941, a crash program was making commercial quantities.[59][67][68]

Large-scale productions of elemental fluorine began during World War II. Germany used high-temperature electrolysis to produce tons of chlorine trifluoride, a compound planned to be used as an incendiary.[69] The Manhattan project in the United States produced even more fluorine for use in uranium separation. Gaseous uranium hexafluoride, was used to separate uranium-235, an important nuclear explosive, from the heavier uranium-238 in centrifuges and diffusion plants.[61] Because uranium hexafluoride releases small quantities of corrosive fluorine, the separation plants were built with special materials. All pipes were coated with nickel, which forms a protective fluoride layer on its surface after exposure to fluorine. Joints and flexible parts were fabricated from Teflon.[68][70]

In 1958, a DuPont research manager in the Teflon business, Bill Gore, left the company because of its unwillingess to develop Teflon as wire-coating insulation. Gore's son Robert found a method for solving the wire-coating problem and the company W. L. Gore and Associates was born. In 1969, Robert Gore developed expanded PTFE membrane which led to the large Gore-tex business in breathable rainwear. The company developed many other uses of PTFE and grew the market for the substance.

In the 1970s and 80s, concerns developed over the role chlorofluorocarbons play in damaging the ozone layer. By 1996, almost all nations had banned chlorofluorocarbon refrigerants and commercial production ceased. Fluorine continued to play a role in refrigeration though: hydrochlorofluorocarbons (HCFCs) and hydrofluorocarbons (HFCs) were developed as replacement refrigerants.[71][72]

Industry and applications

Fluorite mining was estimated in 2003 to be a $550 million industry, extracting 4.5 million tons per year. Two grades of mined fluorite exist: metspar (for iron smelting) is less than 97% CaF2; acidspar (for HF production) is more than 97% CaF2. Approximately half of mined fluorite is used to help molten metal flow, especially in iron smelting. The rest is converted to hydrofluoric acid by reaction with sulfuric acid. HF is primarily used to produce organofluorides and synthetic cryolite. Only about 1% of mined fluorine is ever converted to the element.[51][59]

Uses of inorganic fluorides

About 3 kg of metspar grade fluorite, added directly to the batch, are used for every tonne of steel made. The fluoride ions from CaF2 lower the melt's temperature and viscosity (make the liquid runnier). (The calcium helps remove sulfur and phosphorus, but other additives such as lime do as well.) Metspar is similarly used in cast iron production and for other iron-containing alloys. Fluorite of the acidspar grade is used directly as an additive to ceramics and enamels, glass fibers and clouded glass, and cement, as well as in the outer coating of welding rods.[51]

Acidspar is primarily used for making hydrofluoric acid, which is a chemical intermediate for most fluorine-containing compounds. Significant direct uses of HF include pickling (cleaning) of steel, cracking of alkanes in the petrochemical industry, and etching of glass.[51][73]

One third of HF (one sixth of mined fluorine) is used to make synthetic cryolite (sodium hexafluoroaluminate) and aluminum trifluoride. These compounds are used in the electrolysis of aluminium. About 23 kg are required for every tonne of aluminum. These compounds are also used as a flux for glass.[51][74]

Fluorosilicate is the next most significant inorganic fluoride formed from HF. The material is used for water fluoridation and is also an intermediate for synthetic cryolite. Other inorganic fluorides made in large quantities include cobalt difluoride (for organofluorine synthesis), nickel difluoride (electronics), lithium fluoride (a flux), sodium fluoride (water fluoridation), magnesium difluoride (flux and optical use), potassium fluoride (flux), and ammonium fluoride (various).[51]

Uses of organofluorines

Refrigerant gases

Traditionally chloroflorocarbons (CFCs) have been the predominant fluorinated organic chemical. CFCs (and HFCs and HCFCs) are identified by a system of numbering that explains the amount of fluorine, cholorine, carbon and hydrogen in the molecules. The term Freon has been colloquially used for CFCs and similar halogenated molecules. However, strictly speaking this is just a DuPont brand name; many other producers exist. Brand neutral terminology is to use "R" as the prefix. Prominent CFCs included R-11 (trichlorofluoromethane), R-12 (dichlorodifluoromethane), and R-114 (1,2-dichlorotetrafluoroethane).[51]

Production of CFCs grew strongly through the 1980s, primarily for refrigeration and air conditioning but also for propellants, and solvents. However, since the end use of these materials is banned in most countries, this industry has shrunk dramatically. By the early 21st century, production of CFCs was less than 10% of the mid-80s peak, with remaining use primarily as an intermediate for other chemicals. The banning of CFCs initially depressed the overall demand for fluorite. However, 21st century production of the source mineral has recovered to 1980s levels.[51]

Hydrochlorofluorocarbons (HCFCs) and hydrofluorocarbons (HFCs) now serve as replacements for CFC refrigerants, even though few were commercially manufactured before 1990. Currently more than 90% of fluorine used for organics goes into these two classes (in about equal amounts). Prominent HCFCs include R-22 (chlorodifluoromethane) and R-141b (1,1-dichloro-1-fluoroethane). The main HFC is R-134a (1,1,1,2-tetrafluoroethane).[75][51]

A bromofluoroalkane, "Halon" (bromotrifluoromethane) is still widely used in ship and aircraft gaseous fire suppression systems. Because Halon production has been banned since 1994, systems are dependent on the pre-ban stores and on recycling.[76]

Polymers and coatings

Fluoropolymers were, as of 2008, a $1.4 billion market in the United States alone.[77] PTFE (polytetrafluoroethylene) is 85% of the world's fluoropolymer production. The term Teflon is sometimes used genarically for the substance, but is a DuPont brand name—other PTFE producers exist and also DuPont sometimes uses the Teflon brand for non-PTFE materials. PTFE gets its fluorine without the need for fluorine gas: chloroform (trichloromethane) is treated with HF to make chlorodifluoromethane (R-22, an HFC); this chemical when heated makes ETFE (ethylene tetrafluoride, the starting point for PTFE).[78]

The largest application for PTFE is in electrical insulation. The material has excellent dielectric properties and is not flammable. It is also used extensively in the chemical process industry where corrosion resistance is needed: to coat pipes and in tubing and gaskets. PTFE withstands attack from all chemicals other than alkali metals. Another major use in architectural fabric (PTFE-coated fiberglass cloth used for stadium roofs and such). The major consumer application is in non-stick cookware.[78]

Major PTFE applications
PTFE dielectric separating core and outer metal in a specialty coaxial cable First Teflon frying pan, 1961 The interior of the Tokyo Dome. The roof is PTFE-coated fiberglass, air-supported.[79]


When stretched with a jerk, PTFE films makes a material with fine pores: ePTFE (expanded PTFE) membrane. The term "Gore-tex" is sometimes used generically for this membrane, but that is a specific brand name. W.L. Gore is not the only producer of ePTFE and furthermore "Gore-tex" often refers to more complicated multi-layer membranes or laminated fabrics. ePTFE is used in rainwear, protective apparel and liquids and gas filters. PTFE can also be drawn into fibers which are used in pump packing (seals) and industrial bag house filters for industries with corrosive exhausts.[78]

Other fluoroplymers tend to have similar properties to PTFE—high chemical resistance and good dielectric properties—which leads to use in the chemical process industry and electrical insulation. They are easier to work with (to form into complex shapes), but are more expensive than PTFE and have lower thermal stability. Fluorinated ethylene propylene (FEP) is the second most produced fluoropolymer. Films from two different fluoropolymers serve as glass-replacements in solar cells.[80][81][78]

Elastomers (rubber-like substances) exist that are composed of mixtures of several fluoropolymers. Chemical-resistant O-rings are the primary application. Fluoroelastomers tend to be more stiff and less durable than conventional elastomers. Viton is a prominent example.[78][82]

Fluorinated ionomers are expensive, chemically resistant materials used as membranes in certain electrochemical cells. Nafion, developed in the 1960s, was the first example and remains the most prominent material in the class. The initial Nafion application was as a fuel cell material in spacecraft. Since then, the material has been transforming the 55 million ton per year chloralkali industry; it is replacing hazardous mercury-based cells with membrane cells that are also more energy efficient. While older technology plants continue to run, new plants typically use membrane cells. By 2002, more than a third of the global capacity for the industry was membrane-cell based. Recently, the fuel cell application has reemerged; significant research is being conducted and investments made related to getting proton exchange membrane (PEM) fuel cells into vehicles.[83][84][85][86]

Fluorinated surfactants are small organofluorine molecules, principally used in DWR (durable water repellent). DWR is a finish (thin coating) put on fabrics that makes them lightly rain resistant, that makes water bead. First developed in then 1950s, by 1990 fluorosurfacts were 90% of the DWR industry. In that same year, the market for fluorosurfactants was $100 million in the U.S. textile industry alone. DWR is used with garment fabrics, carpeting, and food packaging. DWR is applied to fabrics by "dip-squeeze-dry" (immersion in DWR-water bath, pressing water out, and then drying). Scotchgard is a prominent brand.[87][88]

Uses of fluorine gas

The largest application for elemental fluorine is the preparation of uranium hexafluoride, which is used in the production of nuclear fuels. To obtain the compound, uranium dioxide is first treated with hydrofluoric acid, to produce uranium tetrafluoride. This compound is then further fluorinated by direct exposure to fluorine gas to make the hexafluoride.[89] Fluorine's monoisotopic natural occurrence makes it useful in uranium enrichment, because uranium hexafluoride molecules will differ in mass only because of mass differences between uranium-235 and uranium-238. These mass differences are used to separate uranium-235 and uranium-238 via diffusion and centrifugation.[51] About 7,000 tons per year of fluorine gas are used for this application.

The second largest application for fluorine gas is sulfur hexafluoride, which is used as an dielectric medium in high voltage switching stations. SF6 gas has a much higher dielectric strength than air. It is extremely inert and, compared to oil-filled switchear, has no PCB (a hazardous chemical). Sulfur hexafluoride is also used in soundproof windows, in the electronics industry, as well as niche medical and military applications. The compound can be made without using fluorine gas, but the reaction between pure sulfur and pure fluorine gas, first developed by Moissan, remains the commercial practice. About 6,000 tons per year of fluorine gas are consumed.[90][51]

Several compounds made from elemental fluorine serve the electronics industry. Tetrafluoromethane,[91] is used for plasma etching in semiconductor manufacturing,[92] flat panel display production, and microelectromechanical systems fabrication.[93][94][89] Nitrogen trifluoride is increasingly used for cleaning equipment at display manufacturing plants. The hexafluorides of rhenium and tungsten are used for chemical vapor deposition (CVD) of thin metal films onto semiconductors. Elemental fluorine, itself, is used sometimes for cleaning equipment.[51]

For making niche organofluorines and fluorine-containing pharmaceuticals, direct fluorination is usually too hard to control. Preparation of intermediate strength fluorinators from fluorine gas solves this problem. The halogen fluorides ClF3, BrF3, and IF5 provide gentler fluorination, with a series of strengths. They are also easier to handle. Sulfur tetrafluoride is used particularly for making fluorinated pharmacueticals. BF3 and SbF3 are not fluorinating agents, but they are formed from fluorine gas and used as catalysts for some organic reactions.[51]

United States and Soviet space scientists in the early 1960s studied elemental fluorine as a possible rocket propellant because of the higher specific impulse generated when fluorine replaced oxygen in combustion. The experiments failed because fluorine proved difficult to handle, and its combustion product (typically hydrogen fluoride) was extremely toxic and corrosive.[95][96][97]

Production of fluorine gas

Commercial producers of fluorine gas continue to use the method of electrolysis pioneerd by Moissan, with some modifications in the cell design. Chemical routes to the elemental form are curiosities that rely on mechanisms to release fluorine gas from compounds previously made using fluorine gas.

Electrolytic synthesis

Several thousand tons of elemental fluorine are produced annually by electrolysis of potassium bifluoride in hydrogen fluoride.[51]

Potassium bifluoride forms spontaneously from potassium fluoride and the hydrogen fluoride:

HF + KF → KHF2

A mixture with the approximate composition KF•2HF melts at 70 °C (158 °F) and is electrolyzed between 70 °C and 130 °C (160–265 °F).[59] Potassium bifluoride increases the electrical conductivity of the solution and provides the bifluoride anion, which releases fluorine at the anode (negative part of the cell). If HF alone is electrolyzed, hydrogen forms at the cathode (positive part of the cell) and the fluoride ions remain in solution. After electrolysis, potassium fluoride remains in solution.[98]

2 HF2 → H2↑ + F2↑ + 2 F

The modern version of the process uses steel containers as cathodes, while blocks of carbon are used as anodes. The carbon electrodes are similar to those used in the electrolysis of aluminium. An earlier version of fluorine production process, by Moissan, uses platinum group metal electrodes and carved fluorite containers. The voltage for the electrolysis is between 8 and 12 volts.[99]

Pure fluorine gas may be stored in steel cylinders where the inside surface is passivated by a metal fluoride layer that resists further attack.[98][59] In the G7 countries, about 17,000 tonnes (17,000,000 kg) of fluorine are produced per year by 11 companies.[89] Fluorine is relatively inexpensive, costing about $5–8 per kilogram when sold as uranium hexafluoride or sulfur hexafluoride. Because of difficulties in storage and handling, the price of pure fluorine gas is much higher.[89]

Chemical routes

In 1986, when preparing for a conference to celebrate the 100th anniversary of the discovery of fluorine, Karl Christe discovered a purely chemical preparation of fluorine gas. It involved the reaction of potassium hexafluoromanganate and antimony pentafluoride at 150 °C, in an atmosphere of hydrogen fluoride:

2 K2MnF6 + 4 SbF5 → 4 KSbF6 + 2 MnF3 + F2

This synthetic route is a rare chemical preparation of elemental fluorine, a reaction not previously thought possible.[100] The manganese(IV) fluoride has to be prepared by reaction with fluorine gas itself,[101] or with krypton difluoride,[102] which is synthesized by reaction with elemental fluorine. This reaction is therefore not an industrially viable way to produce fluorine. Fluorine can by also be chemically synthesized by the reaction of hexafluoronickelate ion NiF2−
6
[49] with the fluorides of the heavier noble gases (krypton[103] and xenon[104]), which can only be produced by a reaction of krypton or xenon with fluorine gas.

Biological aspects

Living organisms

Fluoride is not considered an essential mineral element for mammals and humans, although its role in prevention of tooth decay is well-established.[105] Sodium fluoride, tin(II) fluoride, and, most commonly, sodium monofluorophosphate, are used in toothpaste. These or related compounds, such as fluorosilicates, are added to many municipal water supplies, a process called water fluoridation, which has been controversial since its inception in 1945.[105][106] Small amounts of fluoride may be beneficial for bone strength, but this is an issue only in the formulation of artificial diets.[107]

Biologically synthesized organofluorines have been found in microorganisms and plants,[52] but not in animals.[108] The most common example is fluoroacetate, which is used as a defense against herbivores by at least 40 plants in Australia, Brazil and Africa.[109] Other biologically synthesized organofluorines include ω-fluoro fatty acids, fluoroacetone, and 2-fluorocitrate, all of which are believed to be biosynthesized from the intermediate fluoroacetaldehyde.[108] The enzyme adenosyl-fluoride synthase is capable of biologically synthesizing the carbon–fluorine bond.[110]

Pharmaceuticals, agricultural chemicals, and poisons

Several important pharmaceuticals contain fluorine.[112] Because of the considerable stability of the carbon-fluorine bond, many drugs are fluorinated to prevent their metabolism and prolong their half-lives, allowing for longer times between dosing and activation. For example, an aromatic ring may add to prevent the metabolism of a drug, but this presents a safety problem, because enzymes in the body metabolize some aromatic compounds into poisonous epoxides. Substituting a fluorine into a para position, however, protects the aromatic ring and prevents the epoxide from being produced.[113] Adding fluorine to biologically active organics increases their lipophilicity, because the carbon–fluorine bond is even more hydrophobic than the carbon–hydrogen bond. This effect often increases a drug's bioavailability due to increased cell membrane penetration.[114] Since the carbon–fluorine bond is strong, organofluorides are generally very stable, although the potential of the fluorine to be released as a fluoride leaving group is heavily dependent on the its position in the molecule.

Of drugs that have been commercialized in the past 50 years, 5–15% contain fluorine, and the percentage of currently available fluorine-containing drugs is increasing.[115] For example, fludrocortisone is one of the most common mineralocorticoids, a class of drugs that mimics the actions of aldosterone. The anti-inflammatories dexamethasone and triamcinolone, which are among the most potent of the synthetic corticosteroids class of drugs, contain fluorine.[116] Several inhaled general anesthetic agents, including the most commonly used inhaled agents, also contain fluorine. Examples include sevoflurane, desflurane, and isoflurane, which are hydrofluorocarbon derivatives.[115]

Many SSRI antidepressants are fluorinated organics,[117] including citalopram, escitalopram, fluoxetine, fluvoxamine, and paroxetine. Fluoroquinolones are a commonly used family of broad-spectrum antibiotics.[118] Because biological systems do not metabolize fluorinated molecules easily, fluorinated pharmaceuticals (often antibiotics and antidepressants) are among the major fluorinated organics found in treated city sewage and wastewater.[119]

In addition to pharmaceuticals, an estimated 30% of agrochemical compounds contain fluorine.[120] Because of these, water from agricultural sites contaminates rivers with runoff organofluorines. Synthetic sodium fluoroacetate has been used as an insecticide, especially against cockroaches, and is effective as a bait-poison against mammalian pests.[121] Several other insecticides contain sodium fluoride, which is much less toxic than fluoroacetate.[122]

Isotope use in scanning

Compounds containing fluorine-18, a radioactive isotope that emits positrons, are often used in PET scanning, because its half-life of about 110 minutes is long by positron-emitter standards. One such species is 2-deoxy-2-(18F)fluoro-D-glucose, commonly abbreviated as 18F-FDG.[123][124] In PET imaging, 18F-FDG can be used for the assessment of glucose metabolism in the brain and for imaging tumors in oncology. This radiopharmaceutical is retained by cells and is taken up by tissues with a high need for glucose, such as the brain and most types of malignant tumors.[125] Tomography can thus be used for diagnosis, staging, and monitoring treatment of many types of cancers, particularly in Hodgkin's disease, lung cancer, and breast cancer.

Natural fluorine is monoisotopic, consisting solely of fluorine-19. Fluorine compounds are highly amenable to nuclear magnetic resonance (NMR), because fluorine-19 has a nuclear spin of ½, a high nuclear magnetic moment, and a high magnetogyric ratio, which allows it to make measurements quickly, comparable with a similar effect based on hydrogen-1.[126] Fluorine-19 is commonly used in NMR study of protein structures and conformational changes.[127]

Chronology

Because groundwater contains fluorine ions, organic items such as bone that are buried in soil will absorb those ions over time. As such, it is possible to determine the relative age of an object by comparing the amount of fluoride with another object found in the same area. It is important as a separation technique in intra-site chronological analysis and inter-site comparisons.[128]

However, if no actual age of any object is known, the ages can only be expressed in terms of one of the objects being older or younger than the other. The fluctuating amount of fluoride found in groundwater means the objects being compared must be in the same local area in order for the comparisons to be accurate. This technique is not always reliable, given that not all objects absorb fluorine at the same rates.[129]

Hazards

Elementary state, fluoride ion, and fluoroacetate

Elemental fluorine is a highly toxic, corrosive oxidant, and is extremely reactive to organic material (except for perfluorinated substances) even at very low concentrations and can cause ignition at higher ones.[131] Significant irritation to humans can be caused by concentration of fluorine of 25 ppm; at this and higher concentrations fluorine attacks the eyes, respiratory tract, lungs, liver and kidneys. At a concentration of 100 ppm, human eyes and noses are irritated and seriously damaged.[132]

Soluble fluorides are moderately toxic. In the case of the simple salt sodium fluoride, the lethal dose for most adult humans is estimated at 5 to 10 g, which is equivalent to 32 to 64 mg/kg elemental fluoride/kg body weight.[133][134][135] A toxic dose that may lead to adverse health effects is estimated at 3 to 5 mg/kg of fluoride.[136] The fluoride ion is somewhat toxic, in part because of its ability to form, by equilibration, small amounts of hydrogen fluoride in water. This mobile uncharged species diffuses across cell membranes to attack intracellular calcium. The fluoride ion is readily absorbed by the stomach, intestines and excreted through urine. Urine tests have been used to ascertain rates of excretion in order to set upper limits in exposure to fluoride compounds and associated detrimental health effects.[137] Ingested fluoride initially acts locally on the intestinal mucosa, where it forms hydrofluoric acid in the stomach.[138] Thereafter it binds with calcium and interferes with various enzymes.[138] Excess fluoride consumption can lead to skeletal fluorosis, which currently affects millions of people.[139]

Historically, most cases of fluoride poisoning have been caused by accidental ingestion of insecticides containing inorganic fluoride,[140] or (more rarely) rodenticides containing sodium fluoroacetate ("Compound 1080").[141] Currently, most fluoride poisonings are due to the ingestion of fluoride-containing toothpaste.[138] Malfunction of water fluoridation equipment has occurred several times, including an Alaskan incident, which affected nearly 300 people and killed one person.[142]

Hydrofluoric acid

Hydrofluoric acid is a contact poison, and must be handled with extreme care far beyond that accorded to other mineral acids, even the analogous hydrochloric acid, HCl. Owing to its lesser chemical dissociation in water (remaining a neutral molecule), hydrogen fluoride penetrates tissue more quickly than typical acids. Poisoning can occur readily through exposure of skin or eyes, or when inhaled or swallowed. Symptoms of exposure to hydrofluoric acid may not be immediately evident. Hydrogen fluoride interferes with nerve function, meaning that burns may not initially be painful. Accidental exposures can go unnoticed, delaying treatment and increasing the extent and seriousness of the injury.[143]

Once absorbed into blood through the skin, hydrogen fluoride reacts with blood calcium and may cause cardiac arrest.[144] Formation of insoluble calcium fluoride possibly causes both a fall in calcium serum and the strong pain associated with tissue toxicity.[145] In some cases, exposures can lead to hypocalcemia. Burns with areas larger than 160 cm2 (25 in2) can cause serious systemic toxicity from interference with blood and tissue calcium levels.[146]

Hydrofluoric acid exposure is often treated with calcium gluconate, a source of Ca2+ that binds with the fluoride ions. Hydrogen fluoride chemical burns to the skin can be treated with a water wash and 2.5% calcium gluconate gel[147][148] or special rinsing solutions.[149][150] However, because it is absorbed, medical treatment is necessary; in some cases, amputation may be required.[146]

Environmental concerns

Chlorofluorocarbons and bromofluorocarbons have come under strict environmental regulation due to their long residence times in the atmosphere, and their contribution to ozone depletion. Since it is specifically chlorine and bromine radicals that harm the ozone layer, not fluorine, compounds that do not contain chlorine or bromine but contain only fluorine, carbon, and hydrogen (called hydrofluorocarbons) are not on the United States Environmental Protection Agency list of ozone-depleting substances,[151] and have been widely used as replacements for halocarbons containing chlorine and bromine. Hydrofluorocarbons and perfluorocarbons are greenhouse gases about 4,000 to 10,000 times as potent as carbon dioxide.[152] Sulfur hexafluoride exhibits an even stronger effect, having 20,000 times the global warming potential of carbon dioxide.[152]

Because of the strength of the carbon–fluorine bond, many synthetic fluorocarbons and fluorocarbon-based compounds are persistent in the environment. The fluorosurfactants perfluorooctanesulfonic acid (PFOS) and perfluorooctanoic acid (PFOA), used in waterproofing sprays, and other related chemicals, are persistent global contaminants. PFOS is a persistent organic pollutant and may be harming the health of wildlife.[153] The potential health effects of PFOA to humans are unclear; its tissue distribution in humans is unknown, but studies in rats suggest it is likely to be present primarily in the liver, kidney, and blood, as it is absorbed easily via the gastrointestinal tract in rats. PFOA has been shown not to metabolize in the body, and, unlike chlorinated hydrocarbons, it is neither genotoxic nor lipophilic. It binds to serum albumin and is excreted primarily from the kidney.[154]

Compounds

Fluorine exists in the −1 oxidation state in all compounds except for elemental fluorine, where the atoms are bonded to each other and thus at oxidation state 0. With other atoms, fluorine forms either polar covalent bonds or ionic bonds. Most frequently, covalent bonds involving fluorine atoms are single bonds. Higher bonding can occur; for example boron monofluoride features a triple bond.[155] Fluoride may act as a bridging ligand between two metals in some complex molecules. Molecules containing fluorine may also exhibit hydrogen bonding. Fluorine has a rich chemistry including inorganic compounds formed with hydrogen, metals, non-metals, and even noble gases; as well as a diverse set of organic compounds.[note 2]

Inorganic compounds

Inorganic acids

Unlike other hydrohalic acids, such as hydrochloric acid, hydrofluoric acid is only a weak acid in water solution, with acid dissociation constant (pKa) equal to 3.18.[156] Despite its weakness as an acid in water, hydrogen fluoride is very corrosive, even attacking glass. Because of the basicity of the fluoride ion, soluble fluorides give basic water solutions.

HF's weakness as an acid is seemingly paradoxical considering the polar of the HF bond. Fluoride ions and hydrogen fluoride forms strong hydrogen bonds, unlike the other halides and hydrogen halides. Fluoride ion tends to participate in homoconjugation, forming the highly stable bifluoride ion (HF2-) With less basic solvents such as dry acetic acid, these interactions become less important and hydrofluoric acid then becomes the strongest of the hydrohalic acids.[157][158]

Fluorine forms a family of inorganic acids that contain only a hydrogen, fluorines, and some other element as the central atom. They are generally very strong because of the stable fluorinated anion. One such acid, fluoroantimonic acid (HSbF6), is a "superacid" and the strongest acid known.[159] It has an extremely low pKa of −31.3 and is 20 quintillion (2×1019) times stronger than pure sulfuric acid.[159] Fluoroantimonic acid is so strong that it protonates otherwise inert compounds like hydrocarbons. Hungarian-American chemist George Olah received the 1994 Nobel Prize in chemistry for investigating such reactions.[160]

Metal fluorides

Metal fluorides have similarities with other metal halides and with metal oxides but are more ionic. In general, the other halides (chlorides, bromides, iodides) are more similar to each other than to the fluorides, which often adopt bonding and structures more like oxides.[161] Metal fluorides are solids,[162] liquids,[163] or gases[164] at room temperature.

The fluorides of transition metal elements 25-29
Manganese difluoride
 
Iron trifluoride
 
Cobalt difluoride
 
Nickel difluoride
 
Copper difluoride
 
Ionic metal fluorides

The solubility of ionic fluorides varies greatly but tends to decrease as the charge on the metal ion increases.

The alkali metals form monofluorides. All have the sodium chloride (rock salt) structure and are soluble.[165] Because the fluoride anion is highly basic, many alkali metal fluorides form bifluorides with the formula MHF2. Sodium and potassium bifluorides are significant to the chemical industry.[166] Among other monofluorides, only silver(I)[167] and thallium(I)[168] fluorides are well-characterized. Both are very soluble, unlike the other halides of those metals.

The alkaline earth metals form difluorides that, except for beryllium's, are insoluble.[169] In contrast, the alkaline earth chlorides are readily soluble.[63] Several transition metal difluorides, such as those of copper(II) and nickel(II), are soluble.[63]

Many metals form trifluorides, such as iron, the rare earth elements, and the metals in the aluminium and scandium columns of the periodic table. No trifluoride is soluble in water, but several are soluble in other solvents.[170]

While metal tri- and lower fluorides are all ionic solids, some tetrafluorides are ionic and some molecular. For example, zirconium tetrafluoride is an ionic solid,[171] but germanium tetrafluoride is a molecular gas.[172]

Molecular metal fluorides

Metal penta- and higher fluorides are all molecular and volatile. This behavior contrasts with the corresponding oxides. Oxygen is a weaker oxidant and inherently more likely to form covalent bonds, but it only forms molecules with four metals (manganese heptoxide, technetium heptoxide, ruthenium tetroxide and osmium tetroxide). Fluorine forms molecules with twelve metals because its small size and single charge as an ion allows surrounding metal atoms with more fluorines than oxygen can. (See fluoride volatility.)

Vanadium, niobium, and tantalum lie in a periodic table column that typically reaches +5 as the highest oxidation state. They form pentafluorides as their highest fluoride. Many metals that form hexafluorides also can form pentafluorides.

The metals that make hexafluorides include nine metals in the center of the periodic table (molybdenum, technetium, ruthenium, rhodium, tungsten, rhenium, osmium, iridium, and platinum) along with elements 92–94: uranium, neptunium, and plutonium. At room temperature, tungsten hexafluoride is a gas. Molybdenum hexafluoride and rhenium hexafluoride are liquids. The rest are volatile solids. Metal hexafluorides are oxidants because of their tendency to release fluorines: for example, platinum hexafluoride was the first compound to oxidize molecular oxygen[173] and xenon.[174]

Rhenium is the only metal known to form bonds with seven fluorides, which is the record for number of charged ligands for a charge-neutral metal compound.[175] Rhenium heptafluoride adopts a pentagonal bipyramid molecular geometry. Calculations shows that the currently unknown but perhaps possible compounds osmium heptafluoride and iridium heptafluoride will also have this structure.[176]

Nonmetal fluorides

The nonmetal fluorides are volatile molecules that do not always follow the octet rule of bonding. For instance boron trifluoride has only six electrons around the central boron atom (and thus an incomplete octet). However, the later Period 2 elements form fluorides that follow the octet rule: carbon tetrafluoride,[177] nitrogen trifluoride,[178] and oxygen difluoride.[179][180] Lower periods, however, may form fluorides that are hypervalent molecules, such as phosphorus pentafluoride.[181] The reactivity of such species varies greatly: sulfur hexafluoride is inert, while chlorine fluorides are oxidants, but there are some tendencies and trends based on periodic table columns.

Boron trifluoride is a planar molecule. It is a notable Lewis acid: because of its incomplete octet, it attracts Lewis bases such as ammonia or another fluoride ion which can donate two more electrons.[182][183]

Carbon tetrafluoride and silicon tetrafluoride have tetrahedral structures. CF4 is relatively inert and stable, but SiF4 is a weak acid that is not thermally stable.[177]

Among pnictogens (nitrogen's periodic table column), reactivity and acidity of fluorides increases down the group. Bismuth is an exception—its fluorides are less acidic then those of antimony because bismuth pentafluoride is polymeric,[184] and bismuth trifluoride is ionic (bismuth is a borderline metal).[185] Nitrogen forms no pentafluoride, although tetrafluoroammonium ion, NF+
4
, with nitrogen in the formal oxidation state of +5, is known.[186]

Chalcogens (oxygen's periodic table column) show analagous characteristics: hexafluorides increase in acidity and reactivity down the group.

The halogens form hypervalent compounds. Chlorine[187] and bromine[188] make pentafluorides, both strong fluorinators. Iodine may be fluorinated up to iodine heptafluoride.[189] Astatine is not well-studied, and astatine fluoride has not been produced, even though this should be possible.[190]

The polymeric structure of bismuth pentafluoride

Structure of a (BiF5)n chain
Packing of chains

Noble gas compounds

The noble gases are generally non-reactive because they have fully filled electronic shells, which are extremely stable. Until the 1960s, no chemical bond with a noble gas was known. In 1962, Neil Bartlett used fluorine-containing platinum hexafluoride to react with xenon. He called the compound he prepared xenon hexafluoroplatinate, but since then the product has been revealed to be mixture of different chemicals. Bartlett probably synthesized a mixture of monofluoroxenyl(II) hexafluoroplatinate, [XeF]+[PtF6], monofluoroxenyl(II) undecafluorodiplatinate, [XeF]+[Pt2F11], and trifluorodixenyl(II) hexafluoroplatinate, [Xe2F3]+[PtF6].[191] Bartlett's fluorination of xenon has been called one of the ten most beautiful experiments in the history of chemistry.[192] Later in 1962, xenon was reacted directly with fluorine to form the di- and tetra- fluorides. Since then, chemists have made extensive efforts to form other noble gas fluorides.

Helium and neon form no stable chemical compounds. Helium, forms a temporary helium fluoride hydride that is stable for a few nanoseconds.[193][194] Neon, the least reactive element,[note 3]forms a metastable chemical compound, neon monofluoride, NeF.[196]

Argon forms no binary fluoride, but reacts in extreme conditions with hydrogen fluoride, to form argon fluoride hydride.[197] Argon also forms a short-lived, metastable binary argon monofluoride, ArF, which is used in the argon fluoride laser.[198]

Krypton forms a well-characterized binary noble gas–fluorine compound: krypton difluoride.[199][200][201] Krypton tetrafluoride was reported in 1963,[202] but was subsequently shown to be a mistaken identification.[203][204] A krypton monofluoride radical and cation have been observed at low temperature.[204]

Xenon has the most known noble gas–fluorine compounds. Binary compounds include xenon difluoride, xenon tetrafluoride, and xenon hexafluoride, as well as their derivatives.[199][200][201] Xenon forms several oxyfluorides, such as xenon oxytetrafluoride, XeOF4, by reaction of xenon hexafluoride with water.[205][206]

Radon readily reacts with fluorine to form a solid compound, which is generally thought to be radon difluoride. The exact composition is uncertain—the compound is prone to decompose.[207] Calculations indicate that radon difluoride may be ionic, unlike all other binary noble gas fluorides.[208]

Ununoctium, the last currently known group 18 element, is predicted to form ununoctium difluoride, UuoF2, and ununoctium tetrafluoride, UuoF4, which is likely to have the tetrahedral molecular geometry.[209] However, only a few atoms of ununoctium have been synthesized,[210] and the chemical properties have not been examined.

Highest oxidation states

Elements frequently have their highest oxidation state in the form of a binary fluoride. Several elements show their highest oxidation state only in a few compounds, one of which is the fluoride; and some elements' highest known oxidation state is seen exclusively in a fluoride.[176]

Fluorine was the first element able to oxidize a group 12 element to an oxidation state above +2; making the element's d-electrons participate in bonding.[211] Mercury(IV) fluoride was produced by this reaction, the first ever mercury(IV) compound; its discovery has heated the debate over whether mercury, cadmium, and zinc are transition metals.[212] Another unique oxidation state available for fluorine only is gold(V).[213] It is only known in the hexafluoroaurate(V) ion, which can be synthesized indirectly under extreme conditions, and gold(V) fluoride, which is obtained during hexafluoroaurate(V) decomposition. Because of fluorine's high oxidizing potential it was suggested that gold heptafluoride contained gold(VII),[214] but current calculations show that the claimed AuF7 molecule was AuF5·F2.[215]

Aside from these examples, fluorine is the only element that is known to oxidize the respective elements to palladium(VI),[216] platinum(VI),[217] copper(IV),[218] silver(IV),[219] nickel(IV),[220] iridium(VI),[221] and krypton(II).[201] It is possible that element 113, ununtrium, will be the first boron group element to form a species in the +5 oxidation state, the fluorine-based hexafluoroununtrate(V), UutF
6
;[222] no other ununtrium(V) species is expected.

For groups 1—6, 13, 14, and 16, the highest oxidation states of oxides and fluorides are always equal. Differences are only seen in groups 7—11, mercury, nitrogen, halogens, and the noble gases. Fluorination allows elements to achieve relatively low[note 4] oxidation states that are, however, hard to achieve. For example, no binary oxide is known for krypton, but krypton difluoride is well-studied.[201] At the same time, very high oxidation states are known for oxygen-based species only. For example, ruthenium and xenon octafluorides have not been synthesized, but ruthenium tetroxide and xenon tetroxide are both well-studied.[223] Also, the halogens tend to achieve higher oxidation states as oxides than as fluorides. The main problem that prevents fluorine from forming the highest states in covalent hepta- and octafluorides is that it is hard to attach such a large number of ligands around a single atom; the number of ligands is halved in analogous oxides.[176][note 5]

Organic compounds

Organofluorine compounds are defined by having a carbon–fluorine chemical bond. This bond is the strongest bond in organic chemistry and is very stable.[224] Fluorine replaces hydrogen in hydrocarbons even at room temperature. After the reaction, the molecular size is not changed significantly. Organofluorine compounds are synthesized via both direct reaction with fluorine gas, which can be dangerously reactive, or reaction with fluorinating reagents such as sulfur tetrafluoride.[68] The range of organofluorine compounds is diverse, reflecting the inherent complexity of organic chemistry. A vast number of small molecules exist with varying amounts of fluorine substitution, as well as many polymers—research into particular areas is driven by the commercial value of applications.[68]

Small molecules

Organic compounds with all C-H bonds changed to C-F are called fluorocarbons or perfluorocarbons. They are extremely insoluble in water. The derivatives of alkanes (hydrocarbons with single bonds) have higher density and melting and boiling points and are more thermally and chemically stable. In contrast, fluorinating hydrocarbons which contain double bonds (alkenes) or triple bonds (alkynes) gives rise to very easily attacked molecules. Difluoroacetylene, which is explosive even at low temperatures, is a notable example. Partially fluorinated alkanes exist as well, the hydrofluorocarbons (HFCs).

The Fowler process produces commercial perfluoroalkanes. Hydrocarbon feedstocks are passed over a bed of cobalt trifluoride, which loses fluorine and is reduced to cobalt difluoride. Fluorine reacts with the hydrocarbon producing perfluorocarbon and HF. The cobalt bed is periodically refluorinated to the trifluoride, in a separate step, by reaction with concentrated HF.

Substituting other halogens in combination with fluorine gives rise to chlorofluorocarbons (CFCs) or bromofluorocarbons (BFCs) and the like. (Or if some hydrogen is retained, HCFCs and the like.) Properties depend on the number and identity of the halogen atoms. In general, the boiling points are even more elevated by combination of halogen atoms because the varying size and charge of different halogens allows more intermolecular attractions.[225] As with fluorocarbons, chlorofluorocarbons and bromofluorocarbons are not flammable: they don't have carbon–hydrogen bonds to react and released halides quench flames.[225]

Perfluorinated compounds (as opposed to perfluorocarbons) is the term used for molecules that would be perfluorocarbons—only carbon and fluorine atoms—except for having an extra functional group. The perfluoro parts of the molecule tend to be hydrophobic and slippery. The functional group allows reactions, attachment to solids, solubility, etc. Electrochemical fluorination (ECF), essentially electrolysis in HF, is the commercial production method.

The fluorosurfactants are notable perfluorinated compounds. They have a medium length straight chain of perfluoroalkane terminated by an acid group. This type of arrangement, like a fatty acid, gives rise to a combination of properties. For instance perfluorooctanesulfonic acid (PFOS, formerly the active component in brand "Scotchgard") has eight fluorinated carbons in a row ending with a sulfonic acid group.

For fluorinated organic acids, the large inductive effect of the trifluoromethyl group results in high acid strengths, which may be comparable to mineral acids. In these compounds, the anion's affinity for the acid proton is decreased by the acid's fluorine content, which increases its affinity for the extra electron left when the acidic proton leaves. For example, acetic acid is a weak acid, with pKa equal to 4.76, while its fluorinated derivative, trifluoroacetic acid has pKa of −0.23, giving it 33,000 times greater acid strength.[226]

Polymers

Fluoropolymers are organic polymers ("plastics") containing fluorine. Polytetrafluoroethylene (PTFE, DuPont brand Teflon) is a simple linear chain polymer with the repeating structural unit: –(CF2)–. PTFE has a backbone of carbons single bonded in a long chain, with all side bonds to fluorines. It contains no hydrogens and can be thought of as the perfluoro analog of polyethylene (structural unit: –(CH2)–. PTFE has high chemical and thermal stability, as expected for a perfluorocarbon, much stronger than polyethylene. The non-stick nature of PTFE results from the repulsion of highly charged fluorine atoms in polymeric chains. Its resistance to van der Waals forces makes PTFE the only known surface to which a gecko cannot stick.[227][228]

Several other fluoropolymers exist that are more complicated structurally than PTFE. In general, they have similar properties to Teflon, but are less thermally and chemically stable, while also more processable (lower melting). Three fluoropolymers that contain hydrogen are polyvinylidene fluoride (PVDF, structural unit: –(CF2CH2)–), polyvinyl fluoride (PVF, structural unit: –(CF2CHF)–), and ethylene tetrafluoroethylene (ETFE, structural unit: –(CF2CF2CH2CH2)–).

FEP (fluorinated ethylene propylene) and PFA (perfluoroalkoxy) are similar to PTFE in not containing any hydrogens and contain a PTFE backbone. FEP has branching chains of fluoroalkane. PFA also has branched fluoralkane but with an ether (oxygen) link.

Nafion is a complicated polymer, structurally. It has a PTFE-like backbone, but also contains side chains of perfluoro ether that end in sulfonic acid (SO3H) groups. It has similar chemical stability as PTFE, but because of the acid side chains, is also an ionic conductor, the first ionomer.

See also

Notes

  1. ^ Density of air at 100 kilopascal and 0 °C is 1.2724 grams per liter.[25]
  2. ^ In this article, metalloids are not treated separately from metals and nonmetals, but among elements they are closer to. For example, germanium is treated as a metal, and silicon as a nonmetal. Antimony, bismuth and polonium are included for comparison among nonmetals, even though they are closer to metals chemically than to nonmetals. The noble gases are treated separately from nonmetals and hydrogen is discussed in the Inorganic acids section. P-block period 7 elements have not been studied and thus are not included. This is illustrated by the image to the right: the dark gray elements are metals, the green ones are nonmetals, the light blue ones are the noble gases, the purple one is hydrogen, and the light gray elements have unknown properties.
  3. ^ Helium and neon are the only elements that have no known "stable" compounds; i. e., the compounds that do not decay on specific conditions. Calculations show that helium, unlike neon, may form some compounds that will not decay over time in specific conditions; additionally, clathrates are known for every noble gas but neon.[195]
  4. ^ There is no general line where oxidation states are "relatively low" or "relatively high," they rely on specific elements (and defined only for elements that have highest oxides and fluorides are in different oxidation states); in general, +7 and +8 are high, while +4 and below are low. States +5 and +6 rely on element properties, like atomic radius; for a small nitrogen atom, +5 is what is called "high" here, but for larger palladium and platinum +6 is still "low."
  5. ^ Note that aside from the molecular one, other forms of the highest fluorides are not considered to be possible. Such compounds are extremely unlikely to be ionic due to large sixth, seventh, eight oxidation energies of all elements, which make the reactions that may produce such compounds highly unfavorable. An even higher fluoride of any element seems to be impossible for similar reasons.[176]

References

  1. ^ Wieser, Michael E.; Coplen, Tyler B. (2010). "Atomic weights of the elements 2009 (IUPAC Technical Report)". Pure and Applied Chemistry 83: 359—396. doi:10.1351/PAC-REP-10-09-14. 
  2. ^ a b c Aigueperse et al. 2005, "Fluorine", p. 1.
  3. ^ Aigueperse et al. 2005, "Fluorine", p. 2.
  4. ^ Jarry, Roger L.; Miller, Henry C. (1956). "The Density of Liquid Fluorine between 67 and 103°K.". Journal of the American Chemical Society 78: 1552. doi:10.1021/ja01589a012. 
  5. ^ a b c d Dean 1999, p. 523.
  6. ^ Cengel, Yunus A.; Boles, Michael A. (2002). Thermodynamics: An Engineering Approach (Fourth ed.). McGraw-Hill. p. 824. ISBN 0-07-238332-1. 
  7. ^ a b Dean 1999, p. 942.
  8. ^ a b Compressed Gas Association (1999). Handbook of compressed gases. Springer. p. 365. ISBN 9780412782305. 
  9. ^ a b Allred, A. L. (1961). "Electronegativity values from thermochemical data". Journal of Inorganic and Nuclear Chemistry 17 (3–4): 215–221. doi:10.1016/0022-1902(61)80142-5. 
  10. ^ a b c d e Dean 1999, p. 564.
  11. ^ Robinson, Edward A.; Johnson, Samuel A.; Tang, Ting-Hua; Gillespie, Ronald J. (1997). "Reinterpretation of the Lengths of Bonds to Fluorine in Terms of an Almost Ionic Model". Inorganic Chemistry 36 (14): 3022. doi:10.1021/ic961315b. PMID 11669953. 
  12. ^ Bondi, A. (1964). "Van der Waals Volumes and Radii". Journal of Physical Chemistry 68 (3): 441–51. doi:10.1021/j100785a001. http://pubs.acs.org/doi/pdf/10.1021/j100785a001. 
  13. ^ Young, David A. (1975). Phase Diagrams of the Elements (Report). Springer. p. 10. 
  14. ^ Mackay, Kenneth Malcolm; Mackay, Rosemary Ann; Henderson, W. (2002). Introduction to modern inorganic chemistry (6th ed.). CRC Press. p. 72. ISBN 0748764208. http://books.google.com/books?id=STxHXRR4VKIC&pg=PA72&lpg=PA72&dq=fluorine+diamagnetic&source=bl&ots=EFXyHIc6U0&sig=GKUMwvX_OeiAPlN9LS53wj4WPIY&hl=en&ei=p8D4Tby0B-b00gGfxsm1Cw&sa=X&oi=book_result&ct=result&resnum=7&ved=0CEoQ6AEwBg#v=onepage&q=fluorine%20diamagnetic&f=false. Retrieved 2011-06-15. 
  15. ^ Yaws & Braker 2001, p. 385.
  16. ^ Chiste, V.; Be, M. M. (2006). "F-18". Table de radionucleides. France: Laboratoire National Henri Becquerel. http://www.nucleide.org/DDEP_WG/Nuclides/F-18_tables.pdf. Retrieved 15 June 2011. 
  17. ^ a b Aigueperse et al. 2005, "Fluorine," p. 1.
  18. ^ Blondel, C.; Delsart, C.; Goldfarb, F. (2001). "Electron spectrometry at the μeV level and the electron affinities of Si and F". Journal of Physics B: Atomic, Molecular and Optical Physics (34): L281–L288. doi:10.1088/0953-4075/34/9/101. 
  19. ^ Cordero, Beatriz; Gómez, Verónica; Platero-Prats, Ana E.; Revés, Marc; Echeverría, Jorge; Cremades, Eduard; Barragán, Flavia; Alvarez, Santiago (2008). "Covalent radii revisited". Dalton Translations (21): 2832–38. doi:10.1039/b801115j. 
  20. ^ National Nuclear Data Center. "NuDat 2.1 Database – Fluorine-19". Brookhaven National Laboratory. http://www.nndc.bnl.gov/nudat2/reCenter.jsp?z=9&n=10. Retrieved 2005-09-31. 
  21. ^ a b National Nuclear Data Center. "NuDat 2.1 Database". Brookhaven National Laboratory. http://www.nndc.bnl.gov/nudat2/. Retrieved 2005-09-31. 
  22. ^ National Nuclear Data Center. "NuDat 2.1 Database – Fluorine-18". Brookhaven National Laboratory. http://www.nndc.bnl.gov/nudat2/reCenter.jsp?z=9&n=9. Retrieved 2005-09-31. 
  23. ^ a b Holden, N. E. (2004). "Table of the Isotopes". In D. R. Lide. CRC Handbook of Chemistry and Physics (85th ed.). CRC Press. Section 11. ISBN 978-0849304859. 
  24. ^ a b c Aigueperse et al. 2005, "Fluorine," p. 2.
  25. ^ Shelquist, Richard (2010). "An Introduction to Air Density and Density Altitude Calculations". Shelquist Engineering. http://wahiduddin.net/calc/density_altitude.htm. Retrieved 2011-04-29. 
  26. ^ Burdon, J.; Emson, B.; Edwards, A. J. (1987). "Is fluorine really yellow?". Journal of Fluorine Chemistry 34 (3—4): 471–74. doi:10.1016/S0022-1139(00)85188-X. 
  27. ^ Banks, Ronald Eric; Sharp, D. W. A; Tatlow, J. C. (1986). Fluorine: The First Hundred Years (1886–1986). p. 4. ISBN 9780444750396. http://books.google.com/?id=CB7wAAAAMAAJ. Retrieved 2011-05-02. 
  28. ^ Meyer, Lothar (1969). "Phase transitions in Van der Waal's lattices". In Prigogine, I. Advances in Chemical Physics, volume XVI. John Wiley & Sons, Inc. pp. 364–75. doi:10.1002/9780470143612.ch7. 
  29. ^ Barrett, C. S. (1967). "Argon—Fluorine Phase Diagram". The Journal of Chemical Physics 47 (2): 740–000. doi:10.1063/1.1711946. 
  30. ^ Greenwood & Earnshaw 1998, p. 804.
  31. ^ Moore, John W.; Stanitski, Conrad L.; Jurs, Peter C. (2010). Principles of Chemistry: The Molecular Science. Cengage Learning. p. 156. ISBN 9780495390794. http://books.google.com/?id=ZOm8L9oCwLMC&pg=PA156. Retrieved 2011-05-07. 
  32. ^ Macomber, Roger S. (1996). Organic Chemistry, Volume 1. University Science Books. p. 230. ISBN 0935702903. http://books.google.com/books?id=HyuogOtzoaYC&pg=PA230. Retrieved 2011-07-26. 
  33. ^ Benson, S. W. (1965). "Bond energies". Journal of Chemical Education 42 (9): 502–19. Bibcode 1965JChEd..42..502B. doi:10.1021/ed042p502. 
  34. ^ Young, J. P.; Haire, R. G.; Peterson, J. R.; Ensor, D. D.; Fellow, R. L. (1981). "Chemical consequences of radioactive decay. 2. spectrophotometric study of the ingrowth of berkelium-249 and californium-249 into halides of einsteinium-253". Inorganic Chemistry 20 (11): 3979–83. doi:10.1021/ic50225a076. 
  35. ^ Kratz, J. V. (2003). "Critical evaluation of the chemical properties of the transactinide elements (IUPAC Technical Report)". Pure and Applied Chemistry 75 (1): 103. doi:10.1351/pac200375010103. http://stage.iupac.org/originalWeb/publications/pac/2003/pdf/7501x0103.pdf. Retrieved 2011-05-04. 
  36. ^ Kahn, Bernd (2007). Radioanalytical Chemistry. ISBN 9780387341224. http://books.google.com/?id=9G8fsaXOH9AC&pg=PA356&lpg=PA356. Retrieved 2011-05-04. 
  37. ^ Even, J.; Kratz, W.; Ballof, J.; Buda, R. A.; Eberhardt, K.; Gromm, E.; Hild, D.; Liebe, D. et al. (2008) (PDF). First Transactinide Chemistry Behind TASCA (Report). Gesellschaft für Schwerionenforschung. p. 1. http://www.gsi.de/informationen/wti/library/scientificreport2008/PAPERS/NUSTAR-SHE-12.pdf. Retrieved 2011-04-29. 
  38. ^ Lidin, Molochko & Andreeva 2000, pp. 442–455.
  39. ^ Greenwood & Earnshaw 1998, p. 844.
  40. ^ Aigueperse et al. 2005, "Fluorine," p. 3.
  41. ^ Pitzer, Kenneth S. (1975). "Fluorides of radon and element 118". Journal of the Chemical Society: Chemical communications: 760b—61. doi:10.1039/C3975000760B. 
  42. ^ Khriachtchev, Leonid; Pettersson, Mika; Runeberg, Nino; Lundell, Jan; Räsänen, Markku (2000-08-24). "A stable argon compound". Nature 406 (6798): 874–76. doi:10.1038/35022551. PMID 10972285. http://www.nature.com/nature/journal/v406/n6798/abs/406874a0.html. Retrieved 2011-04-29. 
  43. ^ Lidin, Molochko & Andreeva 2000, p. 252.
  44. ^ Cameron, A. G. W. (1973). "Abundance of the elements in the Solar System". Space Science Review 15: 121–46. Bibcode 1973SSRv...15..121C. doi:10.1007/BF00172440. http://pubs.giss.nasa.gov/docs/1973/1973_Cameron_1.pdf. 
  45. ^ a b Croswell, Ken (September 2003). "Fluorine: An Element–ary Mystery". Sky and Telescope. http://kencroswell.com/fluorine.html. Retrieved 2011-05-03. 
  46. ^ a b Renda, Agostino; Fenner, Yeshe; Gibson, Brad K.; Karakas, Amanda I.; Lattanzio, John C.; Campbell, Simon; Chieffi, Alessandro; Cunha, Katia et al. (2004). "On the origin of fluorine in the Milky Way". Monthly Notices of the Royal Astronomical Society 354 (2): 575–80. arXiv:astro-ph/0410580. Bibcode 2004MNRAS.354..575R. doi:10.1111/j.1365-2966.2004.08215.x. 
  47. ^ Snow, T. P., Jr.; York, D. G. (1981). "The detection of interstellar fluorine in the line of sight toward Delta Scorpii". Astrophysical Journal 247: L39. Bibcode 1981ApJ...247L..39S. doi:10.1086/183585. 
  48. ^ Zhang, Y.; Liu, X.-W. (2005). "Fluorine abundances in planetary nebulae". The Astrophysical Journal 631 (1): L61–63. arXiv:astro-ph/0508339. Bibcode 2005ApJ...631L..61Z. doi:10.1086/497113. http://iopscience.iop.org/1538-4357/631/1/L61/pdf/1538-4357_631_1_L61.pdf. Retrieved 2011-05-03. 
  49. ^ a b Aigueperse et al. 2005, "Fluorine," p. 4.
  50. ^ a b c d e Greenwood & Earnshaw 1998, p. 795.
  51. ^ a b c d e f g h i j k l m n o p Villalba, Gara; Ayres, Robert U.; Schroder, Hans (2008). "Accounting for fluorine: production, use, and loss". Journal of Industrial Ecology 11: 85–101. doi:10.1162/jiec.2007.1075. 
  52. ^ a b Gribble, Gordon W. (2002). "Naturally occurring organofluorines". The Handbook of Environmental Chemistry 3N: 121–36. doi:10.1007/10721878_5. 
  53. ^ Greenwood & Earnshaw 1998, p. 790.
  54. ^ Senning, Alexander (2007). Elsevier's Dictionary of Chemoetymology: The Whies and Whences of Chemical Nomenclature and Terminology. Elsevier. p. 149. ISBN 9780444522399. http://books.google.com/?id=Fl4sdCYrq3cC&pg=PT149. Retrieved 2011-05-07. 
  55. ^ Agricola, Georgius (2003) [1912]. De Re Metallica. trans. Hoover, Herbert Clark; Hoover, Henry Lou. p. 381. ISBN 9780766131972. http://books.google.com/?id=oWoChuYV2GUC&pg=PA381. 
  56. ^ Greenwood & Earnshaw 1998, p. 109.
  57. ^ a b c Asimov, Isaac (1966). The Noble Gases. ISBN 978-0465051298
  58. ^ Fluorum: Fluorine vanderkrogt.net. Last accessed 26 Dec 2011.
  59. ^ a b c d e f g Kirsch, Peer (2004). "Fluorine". Modern Fluoroorganic Chemistry: Synthesis, reactivity, applications. pp. 3–10. ISBN 9783527306916. http://books.google.com/?id=ycSf3EK9i2gC&pg=PA5. Retrieved 2011-05-07. 
  60. ^ (French)Ampere, André-Marie (1816). "Suite d'une Classification naturelle pour les Corps simples" (in French). Annales de chimie et de physique: 1–5. http://books.google.com/?id=4jEFAAAAQAAJ&pg=RA1-PA5. Retrieved 2011-05-07. 
  61. ^ a b Peter, Meiers (2010). "Discovery of Fluorine". Fluoride History. http://www.fluoride-history.de/fluorine.htm. Retrieved 2011-01-17. 
  62. ^ a b Weeks, Mary Elvira (1932). "The discovery of the elements. XVII. The halogen family". Journal of Chemical Education 9 (11): 1915–39. Bibcode 1932JChEd...9.1915W. doi:10.1021/ed009p1915. 
  63. ^ a b c Storer, Frank Humphreys (1864). First Outlines of a Dictionary of Solubilities of Chemical Substances. Cambridge. pp. 278–80. ISBN 978-1176622562. 
  64. ^ Moissan, Henry (1886). "Action d'un courant électrique sur l'acide fluorhydrique anhydre" (in French). Comptes rendus hebdomadaires des séances de l'Académie des sciences 102: 1543–44. http://gallica.bnf.fr/ark:/12148/bpt6k3058f/f1541.chemindefer. Retrieved 2011-05-07. 
  65. ^ "The Nobel Prize in Chemistry 1906". Nobelprize.org. http://nobelprize.org/nobel_prizes/chemistry/laureates/1906/. Retrieved 2009-07-07. 
  66. ^ Greenwood & Earnshaw 1998, pp. 789–791.
  67. ^ a b Hounshell, David A.; Smith, John Kenly (1988). Science and Corporate Strategy: DuPont R&D, 1902–1980. Cambridge University Press. pp. 147, 156–57, 482–484. ISBN 0521327679. 
  68. ^ a b c d e Okazoe, Takashi (2009). "Overview on the history of organofluorine chemistry from the viewpoint of material industry". Proceedings of the Japan Academy, Series B 85 (8): 276–89. Bibcode 2009PJAB...85..276O. doi:10.2183/pjab.85.276. 
  69. ^ Leech, H. R. (1949). "Laboratory and technical production of fluorine and its compounds". Quarterly Reviews, Chemical Society 3: 22–35. doi:10.1039/QR9490300022. 
  70. ^ Hendricks, James O. (1953). "Industrial fluorochemicals". Industrial & Engineering Chemistry 45: 99–105. doi:10.1021/ie50517a034. 
  71. ^ US EPA (2008-09-03). "Ozone Depletion Glossary". http://www.epa.gov/ozone/defns.html#hcfc. Retrieved 2008-09-03. 
  72. ^ "Brief Questions and Answers on Ozone Depletion | Ozone Layer Protection | US EPA". Epa.gov. 2006-06-28. http://www.epa.gov/ozone/science/q_a.html. Retrieved 2011-11-08. 
  73. ^ Pulker, Hans K. (1999). Coatings on Glass. ISBN 9780444501035. http://books.google.com/?id=ho2pYhOiEMcC&pg=PA3. 
  74. ^ Harper, Charles A. (2001). Handbook of Ceramics, Glasses, and Diamonds. p. A26. ISBN 9780070267121. http://books.google.com/?id=_T9dX14rz64C&pg=SL1-PA26. Retrieved 2011-05-07. 
  75. ^ Smart, Bruce E.; Tatlow, J. C. (1994). "Aliphatic Fluorides Go Commercial". Organofluorine Chemistry: Principles and Commercial Applications. pp. 12–15. ISBN 9780306446108. http://books.google.com/?id=TRMRDXHWX7EC&pg=PA12. 
  76. ^ "Executive summary". Fire Suppression Substitutes and Alternatives to Halon for U.S. Navy Applications. Washington, D.C.: National Academies Press. 1997. p. 1. ISBN 9780309074926. http://www.nap.edu/openbook.php?record_id=5744&page=1. 
  77. ^ "Fluoropolymers to 2013 - Demand and Sales Forecasts, Market Share, Market Size, Market Leaders (abstract for market report)". Freedonia Group. http://www.freedoniagroup.com/Fluoropolymers.html. Retrieved 2011-12-31. 
  78. ^ a b c d e Martin, John Wilson (2007). Concise encyclopedia of the structure of materials. Oxford: Elsevier. pp. 187-194. ISBN 978-0-08-045127-5. http://books.google.com/books?id=xv420pEC2qMC&pg=PA189&dq=what+are+the+major+uses+of+PTFE&hl=en&sa=X&ei=GY7_TpW3Gern0QGP5cWtAg&ved=0CEYQ6AEwAA#v=onepage&q=what%20are%20the%20major%20uses%20of%20PTFE&f=false. 
  79. ^ Nakagawa, Ulara. "15 sights that make Tokyo so fascinating". CNN. http://www.cnngo.com/tokyo/life/15-reasons-why-tokyo-rules-583723. Retrieved 2011-12-31. 
  80. ^ Bhiwankar, Nikhil (October/November, 2011). "Weathering the Storm: Fluoropolymer films protect solar modules and ensure performance". Altenergy,com. http://www.altenergymag.com/emagazine/2011/10/weathering-the-storm-fluoropolymer-films-protect-solar-modules-and-ensure-performance/1800. Retrieved 2011-12-31. 
  81. ^ DeBergalis, Michael (2004). "Fluoropolymer films in the photovoltaic industry". Journal of Fluorine Chemistry 125: 1255-1257. http://www2.dupont.com/Photovoltaics/en_US/assets/downloads/pdf/DeBergalis_flpr_films_wp.pdf. 
  82. ^ Klingender, Robert C. (Boca Raton). Handbook of specialty elastomers. 2008: CRC Press. pp. 133-154. ISBN 978-1-57444-676-0. http://books.google.com/books?id=AUNa7lZseLAC&pg=PA230&dq=viton+elastomer&hl=en&sa=X&ei=grwAT8W4NKX20gH13rC0Ag&ved=0CFMQ6AEwAg#v=onepage&q=viton&f=false. 
  83. ^ Mauritz, Kenneth A.; Moore, Robert B. (2004). "State of understanding of Nafion". Chemical Reviews 104 (10): 4535–86. doi:10.1021/cr0207123. PMID 15669162. 
  84. ^ Grot, Walter (2011). Fluorinated Ionomers. Oxford: Elsevier. pp. 1-10. ISBN 978-1-4377-4457-6. http://books.google.com/books?id=E8H1Hwd5GXUC&pg=PA6&dq=nafion+chloralkali&hl=en&sa=X&ei=4sYAT96WLuj10gHs8ICJAg&ved=0CHMQ6AEwCQ#v=onepage&q=nafion%20chloralkali&f=false. 
  85. ^ Ramkumar, Jayshreee (2012). "Nafion Persulphonate Membrane: Unique Properties and Various Applications". In Banerjee, S.. Functional Materials: Preparation, Processing and Applications. London: Elsevier. pp. 549-578. ISBN 978-0-12-385142-0. http://books.google.com/books?id=ep7U5G4O3mQC&pg=PA567&dq=nafion+chloralkali&hl=en&sa=X&ei=4sYAT96WLuj10gHs8ICJAg&ved=0CD4Q6AEwAA#v=onepage&q=nafion%20&f=false. 
  86. ^ Burney, H.S. (1999). "Past, Present, and Future of the Chlor-Alkali Industry". Chlor-alkali and chlorate technology: R.B. MacMullin memorial symposium. Pennington, New Jersey: The Electrochemical Society. pp. 105-126. ISBN 1-56777-244-3. http://books.google.com/books?id=rtm_bjQd8SUC&pg=PA111&lpg=PA111&dq=what+percent+of+the+chlor+alkali+capacity+is+membrane+cell?&source=bl&ots=mLy4epWbFa&sig=5DkLLjuODIXLlvj-xICPq_zI2y8&hl=en&sa=X&ei=i9gAT_n1IqHH0AGhqIWxAg&ved=0CCMQ6AEwAQ#v=onepage&q=what%20percent%20of%20the%20chlor%20alkali%20capacity%20is%20membrane%20cell%3F&f=false. 
  87. ^ Kissa, Erik (2001). Fluorinated surfactants and repellents. New York: Marcel Dekker. pp. 516-551. ISBN 0-8247-0472-X. http://books.google.com/books?id=iAmE8v3bFnUC&pg=PA522&dq=durable+water+repellent&hl=en&sa=X&ei=suEAT9nWLMnb0QG607mwCQ&ved=0CF4Q6AEwBQ#v=onepage&q=durable%20water%20repellent&f=false. 
  88. ^ Ullman, Fritz (2008). Ullmann's fibers: Textile and dyeing technologies, high performance and optical fibers, Volume 2. Germany: Wiley-VCH. pp. 543-547. ISBN 978-3-527-31772-1. http://books.google.com/books?id=OKBKwGV5rSsC&pg=PA543&dq=durable+water+repellent&hl=en&sa=X&ei=suEAT9nWLMnb0QG607mwCQ&ved=0CGMQ6AEwBg#v=onepage&q=durable%20water%20repellent&f=false. 
  89. ^ a b c d Aigueperse et al. 2005, "Fluorine," p. 12.
  90. ^ Aigueperse et al. 2005, "Fluorine Compounds, Inorganic," p. 35.
  91. ^ Aigueperse et al. 2005, "Fluorine," p. 11.
  92. ^ El-Kareh, Badih (1994). "Fluorine –Based Plasmas". Fundamentals of Semiconductor Processing Technology. p. 317. ISBN 9780792395348. http://books.google.com/?id=Wg1SnT6Cb0QC&pg=PA317. Retrieved 2011-05-07. 
  93. ^ Arana, Leonel R.; de Mas, Nuria; Schmidt, Aleksander J.; Franz, Martin A.; Jensen, Schmidt F.; Jensen, Klaus F. (2007). "Isotropic etching of silicon in fluorine gas for MEMS micromachining". Journal Micromechanical Microenergy 17 (2): 384. Bibcode 2007JMiMi..17..384A. doi:10.1088/0960-1317/17/2/026. 
  94. ^ Bahreyni, Behraad (2008). "Vapour Phase Etching". Fabrication and Design of Resonant Microdevices. pp. 27–28. ISBN 9780815515777. http://books.google.com/?id=GyyA5b_n1bUC&pg=PA27. Retrieved 2011-05-07. 
  95. ^ Krieger, F. J. (1960). The Russian Literature on Rocket Propellant (Report). The Rand Corporation. p. 17. http://www.rand.org/pubs/papers/2006/P1954.pdf. Retrieved 2011-02-31. 
  96. ^ Sutton, Oscar; Biblarz (2010). "Liquid Oxidizers". Rocket Propulsion Elements. p. 256. ISBN 9780470080245. http://books.google.com/?id=1Sf6eV6CgtEC&pg=PA256. Retrieved 2011-05-07. 
  97. ^ Tormey, John (1957). "Liquid rocket propellants — Is there an energy limit?". Industrial & Engineering Chemistry 49 (9): 1339–43. doi:10.1021/ie50573a021. 
  98. ^ a b Aigueperse et al. 2005, "Fluorine," pp. 4–5.
  99. ^ Aigueperse et al. 2005, "Fluorine," p. 6.
  100. ^ Christe, K. (1986). "Chemical synthesis of elemental fluorine". Inorganic Chemistry 25 (21): 3721–24. doi:10.1021/ic00241a001. 
  101. ^ Mazej, Z. (2002). "Room temperature syntheses of MnF3, MnF4 and hexafluoromanganete(IV) salts of alkali cations". Journal of Fluorine Chemistry 114 (1): 75–80. doi:10.1016/S0022-1139(01)00566-8. 
  102. ^ Lutar, Karel; Jesih, Adolf; Žemva, Boris (1988). "KrF2/MnF4 adducts from KrF2/MnF2 interaction in HF as a route to high purity MnF4". Polyhedron 7 (13): 1217–19. doi:10.1016/S0277-5387(00)81212-7. 
  103. ^ Lehmann, J. F.; Mercier, H. L. N. P. A.; Schrobilgen, G. J. (2002). "The chemistry of krypton". Coordination Chemistry Reviews 233–234: 1–39. doi:10.1016/S0010-8545(02)00202-3. 
  104. ^ Tius, M. A. (1995). "Xenon difluoride in synthesis". Tetrahedron 51 (24): 6605–34. doi:10.1016/0040-4020(95)00362-C. 
  105. ^ a b Olivares, M.; Uauy, R. (2004). "Essential Nutrients in Drinking Water (Draft)". WHO. http://www.who.int/water_sanitation_health/dwq/en/nutoverview.pdf. Retrieved 2008-12-30. 
  106. ^ Pizzo, G.; Piscopo, M. R.; Pizzo, I.; Giuliana, G. (2007). "Community water fluoridation and caries prevention: a critical review". Clinical Oral Investigations 11 (3): 189–93. doi:10.1007/s00784-007-0111-6. PMID 17333303. 
  107. ^ Nielsen, Forrest H. (2009). "Micronutrients in parenteral nutrition: Boron, silicon, and fluoride". Gastroenterology 137 (5 Suppl): S55–S60. doi:10.1053/j.gastro.2009.07.072. PMID 19874950. 
  108. ^ a b Murphy, C.; Schaffrath, C.; O'Hagan, D. (2003). "Fluorinated natural products: the biosynthesis of fluoroacetate and 4-fluorothreonine in Streptomyces cattleya". Chemosphere 52 (2): 455–61. doi:10.1016/S0045-6535(03)00191-7. PMID 12738270. 
  109. ^ Proudfoot, A. T.; Bradberry, S. M.; Vale, J. A. (2006). "Sodium fluoroacetate poisoning". Toxicological Reviews 25 (4): 213–19. doi:10.2165/00139709-200625040-00002. PMID 17288493. 
  110. ^ O'Hagan, D.; Schaffrath, C.; Cobb, S. L.; Hamilton, J. T.; Murphy, C. D. (2002). "Biochemistry: biosynthesis of an organofluorine molecule". Nature 416 (6878): 279. doi:10.1038/416279a. PMID 11907567. 
  111. ^ Seth, S. D.; Seth, Vimlesh. Textbook of Pharmacology (3rd ed.). Elsevier India. p. VE-40. ISBN 9788131211588. 
  112. ^ National Research Council (U.S.). Committee on Fluoride in Drinking Water (2006). Fluoride in Drinking Water: A Scientific Review of EPA's Standards. National Academies Press. p. 49. ISBN 978-0309101288. 
  113. ^ Rentmeister, A.; Arnold, F. H.; Fasan, R. (January 2009). "Chemo-enzymatic fluorination of unactivated organic compounds". Nature Chemical Biology 5 (1): 26–28. doi:10.1038/nchembio.128. PMC 2713661. PMID 19011638. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2713661. 
  114. ^ "Fluorine - A Vital Element in the Medicine Chest" (PDF). Pharmaceutical Chemistry. 2005. pp. 1–4. http://www.halocarbon.com/halocarbon_media/swinson_109.pdf. Retrieved 2010-08-26. 
  115. ^ a b Filler, R.; Saha, R. (August 2009). "Fluorine in medicinal chemistry: a century of progress and a 60-year retrospective of selected highlights". Future Medicinal Chemistry 1 (5): 777–91. doi:10.4155/fmc.09.65. PMID 21426080. 
  116. ^ Hagmann, W.K. (August 2008). "The many roles for fluorine in medicinal chemistry". Journal of Medicinal Chemistry 51 (15): 4359–69. doi:10.1021/jm800219f. PMID 18570365. 
  117. ^ Mitchell, Siobhan E.; Triggle, D. J. (2004). Antidepressants. Infobase Publishing. pp. 37–39. ISBN 9781438101927. 
  118. ^ Nelson, J.M.; Chiller, T. M.; Powers, J. H.; Angulo, F. J. (2007). "Fluoroquinolone-resistant Campylobacter species and the withdrawal of fluoroquinolones from use in poultry: a public health success story". Clinical Infectious Diseases 44 (7): 977–80. doi:10.1086/512369. PMID 17342653. 
  119. ^ Lietz, A. C.; Meyer, Michael T. (2006). Evaluation of Emerging Contaminants of Concern at the South District Wastewater Treatment Plant Based on Seasonal Sampling Events, Miami-Dade Country, Florida, 2004 (Report). U.S. Geological Survey Scientific Investigations. pp. 7–8. 
  120. ^ "Fluorine's Treasure Trove". ICIS news. 2006. http://www.icis.com/Articles/2006/09/30/2016413/fluorines-treasure-trove.html. Retrieved 2011-02-20. 
  121. ^ Eisler, Ronald (February 1995). Biological Report 27: Sodium monofluoroacetate (1080) Hazards to Fish, Wildlife and Invertebrates: A Synoptic Review (Report). Patuxent Environmental Science Center (U.S. National Biological Service). http://www.pwrc.usgs.gov/infobase/eisler/CHR_30_Sodium_monofluoroacetate.pdf. Retrieved 2011-06-05. 
  122. ^ "Class I Ozone-Depleting Substances". Sodium fluoride – pesticidal uses. Scorecard. http://scorecard.goodguide.com/chemical-profiles/pesticides.tcl?edf_substance_id=7681-49-4. Retrieved 2011-02-20. 
  123. ^ Georges, Audi (2003). "The NUBASE evaluation of nuclear and decay properties". Nuclear Physics A (Atomic Mass Data Center) 729: 3–128. Bibcode 2003NuPhA.729....3A. doi:10.1016/j.nuclphysa.2003.11.001. 
  124. ^ Schmitz, A.; Kälicke, T.; Willkomm, P.; Grünwald, F.; Kandyba, J.; Schmitt, O. (2000). "Use of fluorine-18 fluoro-2-deoxy-D-glucose positron emission tomography in assessing the process of tuberculous spondylitis". Journal of spinal disorders 13 (6): 541–44. doi:10.1097/00002517-200012000-00016. PMID 11132989. 
  125. ^ Bustamante, Ernesto; Pedersen, Peter L. (1977). "High aerobic glycolysis of rat hepatoma cells in culture: Role of mitochondrial hexokinase". Proceedings of the National Academy of Sciences 74 (9): 3735–39. Bibcode 1977PNAS...74.3735B. doi:10.1073/pnas.74.9.3735. PMC 431708. PMID 198801. http://www.pnas.org/content/74/9/3735.full.pdf+html. Retrieved 2011-05-07. 
  126. ^ Nelson, J. H. (2003). Nuclear Magnetic Resonance Spectroscopy. Prentice Hall. pp. 129–39. ISBN 0-13-033451-0. http://books.google.com/?ei=5mlhTambF4fHswbu3_S1CA. Retrieved 2011-05-07. 
  127. ^ Danielson, Mark A.; Falke, Joseph J. (1996). "Use of 19F NMR to probe protein structure and conformational changes". Annual Review of Biophysics and Biomolecular Structure 25: 163–95. doi:10.1146/annurev.bb.25.060196.001115. PMC 2899692. PMID 8800468. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2899692. 
  128. ^ Göksu Murphy, H. Y.; Oberhofer, Martin; Regulla, D. F. (1991). Scientific Dating Methods. Kluver Academic Publishers. p. 267. ISBN 978-0-7923-1461-5. http://books.google.com/?id=maZIlXibg5YC&pg=PA267. Retrieved 2011-05-07. 
  129. ^ Reiche, I. (2006). Fluorine and its relevance for archaeological studies. 2. pp. 253–83. doi:10.1016/S1872-0358(06)02008-2. 
  130. ^ "UN/NA 1045 (United Nations/North America Fluorine Data Sheet)". Cameo Chemicals. National Oceanic and Atmospheric Administration. http://cameochemicals.noaa.gov/unna/1045. Retrieved 17 June 2011. 
  131. ^ Aigueperse et al. 2005, "Fluorine Compounds, Organic," pp. 3–4.
  132. ^ Keplinger, M. L.; Suissa, L. W. (1968). "Toxicity of fluorine short-term inhalation". American Industrial Hygiene Association Journal 29 (1): 10–18. doi:10.1080/00028896809342975. PMID 5667185. 
  133. ^ Gosselin, R. E.; Smith, R. P.; Hodge, H. C. (1984). Clinical Toxicology of Commercial Products. Williams & Wilkins. pp. III–185–93. ISBN 0683036327. 
  134. ^ Baselt, R.C. (2008). Disposition of Toxic Drugs and Chemicals in Man. Biomedical Publications. pp. 636–40. ISBN 0962652370. 
  135. ^ IPCS (2002). Environmental Health Criteria 227 (Fluoride). International Programme on Chemical Safety, World Health Organization. p. 100. ISBN 9241572272. 
  136. ^ Nochimson, Geofrey (11 May 2011). "Fluoride Toxicity". emedicine. http://emedicine.medscape.com/article/814774-overview. Retrieved 2011-05-19. 
  137. ^ Baez, Ramon J.; Baez, Martha X.; Marthaler, Thomas M. (2000). "Urinary fluoride excretion by children 4–6 years old in a south Texas community". Revista Panamericana de Salud Pública 7 (4): 242–248. doi:10.1590/S1020-49892000000400005. 
  138. ^ a b c Nochimson, G. (2008). "Toxicity, Fluoride". eMedicine. http://emedicine.medscape.com/article/814774-overview. Retrieved 2008-12-28. 
  139. ^ Reddy, D. R. (2009). "Neurology of endemic skeletal fluorosis". Neurology India 57 (1): 7–12. doi:10.4103/0028-3886.48793. PMID 19305069. http://www.neurologyindia.com/article.asp?issn=0028-3886;year=2009;volume=57;issue=1;spage=7;epage=12;aulast=Reddy. 
  140. ^ Augenstein, W. L.; Spoerke, D. G.; Kulig, K. W.; Hall, A. H.; Hall, P. K.; Riggs, B. S.; El Saadi, M.; Rumack, B. H. (1991). "Fluoride ingestion in children: a review of 87 cases". Pediatrics 88 (5): 907–12. PMID 1945630. http://pediatrics.aappublications.org/cgi/content/abstract/88/5/907. Retrieved 2011-05-02. 
  141. ^ Green, W. (July 2004). "The Use of 1080 for Pest Control" (PDF). The Animal Health Board and The Department of Conservation. pp. 10–55. http://www.doc.govt.nz/upload/documents/conservation/threats-and-impacts/animal-pests/use-of-1080-04.pdf. Retrieved 2008-12-16. 
  142. ^ Gessner, Bradford D.; Beller, Michael; Middaugh, John P.; Whitford, Gary M. (1994-01-13). "Acute fluoride poisoning from a public water system". New England Journal of Medicine 330 (2): 95–99. doi:10.1056/NEJM199401133300203. PMID 8259189. http://content.nejm.org/cgi/content/full/330/2/95. Retrieved 2011-05-02. 
  143. ^ Yamashita, Masatomo; Yamashita, Mamoru; Suzuki, Mikio; Hirai, Hiroyasu; Kajigaya, Hiroshi (2001). "Ionophoretic delivery of calcium for experimental hydrofluoric acid burns". Critical Care Medicine 29 (8): 1575–78. doi:10.1097/00003246-200108000-00013. PMID 11505130. 
  144. ^ Blodgett, David W.; Suruda, Anthony J.; Crouch, Barbara Insley (2001). "Fatal unintentional occupational poisonings by hydrofluoric acid in the U.S". American Journal of Industrial Medicine 40 (2): 215–20. doi:10.1002/ajim.1090. PMID 11494350. 
  145. ^ Goldfrank's Manual of Toxicologic Emergencies. New York: McGraw-Hill Professional. 2007. p. 1333. ISBN 978-0071443104. http://books.google.com/books?id=gWvDBZrTA8MC. Retrieved 2011-05-03. 
  146. ^ a b "Recommended Medical Treatment for Hydrofluoric Acid Exposure" (PDF). Honeywell Specialty Materials. pp. 2–12. http://www51.honeywell.com/sm/hfacid/common/documents/HF_medical_book.pdf. Retrieved 2009-05-06. 
  147. ^ El Saadi, M. S.; Hall, A. H.; Hall, P. K.; Riggs, B. S.; Augenstein, W. L.; Rumack, B. H. (1989). "Hydrofluoric acid dermal exposure". Veterinary and human toxicology 31 (3): 243–47. PMID 2741315. 
  148. ^ Roblin, Isabelle; Urban, Martine; Flicoteau, Domitille; Martin, Chantel; Pradeau, Dominique (2006). "Topical treatment of experimental hydrofluoric acid skin burns by 2.5% calcium gluconate". Journal of Burn Care & Research 27 (6): 889–94. doi:10.1097/01.BCR.0000245767.54278.09. PMID 17091088. 
  149. ^ Hultén, Peter; Höjer, J.; Ludwigs, U.; Janson, A. (2004). "Hexafluorine vs. standard decontamination to reduce systemic toxicity after dermal exposure to hydrofluoric acid". Journal of Toxicology — Clinical Toxicology 42 (4): 355–361. doi:10.1081/CLT-120039541. PMID 15461243. 
  150. ^ "News & views". Chemical Health and Safety 12 (5): 35–37. 2005. doi:10.1016/j.chs.2005.07.007. 
  151. ^ "Class I Ozone-Depleting Substances". Ozone Depletion. U.S. Environmental Protection Agency. 2010-08-19. http://www.epa.gov/ozone/science/ods/classone.html. Retrieved 2011-02-20. 
  152. ^ a b Intergovernmental Panel on Climate Change. "Changes in Atmospheric Constituents and in Radiative Forcing." (PDF). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate. Cambridge University. p. Table 2.14, Chap. 2, p. 212. ISBN 9780521705967. http://www.ipcc.ch/pdf/assessment-report/ar4/wg1/ar4-wg1-chapter2.pdf. 
  153. ^ Betts, Kellyn S. (1 May 2007). "Perfluoroalkyl acids: What is the evidence telling us?". Environmental Health Perspectives 115 (5): A250–56. doi:10.1289/ehp.115-a250. PMC 1867999. PMID 17520044. http://ehp03.niehs.nih.gov/article/info%3Adoi%2F10.1289%2Fehp.115-a250. Retrieved 5 September 2011. 
  154. ^ Steenland, Kyle; Fletcher, Tony; Savitz, David A. (2010). "Epidemiologic evidence on the health effects of perfluorooctanoic acid (PFOA)". Environmental Health Perspectives 118 (8): 1100. doi:10.1289/ehp.0901827. PMC 2920088. PMID 20423814. http://ehp03.niehs.nih.gov/article/fetchArticle.action?articleURI=info%3Adoi%2F10.1289%2Fehp.0901827. Retrieved 2011-05-03. 
  155. ^ Allan McQuarrie, Donald; Simon, John Douglas (1997). Physical Chemistry: A Molecular Approach. University Science Books. p. 365. ISBN 0-935702-99-7. 
  156. ^ Herring, F. Geoffrey; Petrucci, Ralph H.; Harwood, William (2002). General Chemistry: Principles and Modern Applications (8th ed.). Prentice Hall. p. 678. ISBN 0-13-014329-4. 
  157. ^ Jolly, W. L. (1984). Modern Inorganic Chemistry. McGraw-Hill. p. 203. ISBN 0-07-032760-2. 
  158. ^ Cotton, F. A.; Wilkinson, G. (1988). Advanced Inorganic Chemistry. John Wiley and Sons. p. 109. ISBN 0-471-84997-9. 
  159. ^ a b Olah, George A. (2005). "Crossing conventional boundaries in half a century of research". Journal of Organic Chemistry 70 (7): 2413–29. doi:10.1021/jo040285o. PMID 15787527. 
  160. ^ "The Nobel Prize in Chemistry 1994". The Nobel Foundation. http://nobelprize.org/nobel_prizes/chemistry/laureates/1994/index.html. Retrieved 2008-12-22. 
  161. ^ Greenwood & Earnshaw 1998, p. 819.
  162. ^ Lide, David R. (2009). Handbook of Chemistry and Physics (90 ed.). CRC Press. pp. 4–81. ISBN 978-1-4200-9084-0. 
  163. ^ Bernhardt, N. A.; Bishop, H. W.; Brusie, J. P.; U.S. Atomic Energy Commission (1947). The Preparation of Molybdenum Hexafluoride (Report). pp. 1–2. 
  164. ^ Neikov, Oleg D.; Naboychenko, Stanislav; Gopienko, Victor G.; Frishberg, Irina V. (2008). Handbook of Non-Ferrous Metal Powders: Technologies and Applications. Elsevier. p. 440. ISBN 9781856174220. 
  165. ^ Aigueperse et al. 2005, "Fluorine Compounds, Inorganic," pp. 25–27.
  166. ^ Aigueperse et al. 2005, "Fluorine Compounds, Inorganic," pp. 26–27.
  167. ^ Milne, George W. A. (2005). Gardner's commercially important chemicals. John Wiley and Sons. p. 553. ISBN 978-0-471-73518-2. 
  168. ^ Arora, M. G. (2003). P-block Elements. Anmol Publications. p. 35. ISBN 8174885633. 
  169. ^ U.S. Department of Health and Human Services (1998). Barry Leonard. ed. Eighth Annual Report on Carcinogens: 1998 Summary. Institut d'Estudis Catalans. p. 63. ISBN 9780788183966. 
  170. ^ Sobolev, Boris Petrovich (2001). The Rare Earth Trifluorides: Introduction to materials science of multicomponent metal fluoride crystals. Institut d'Estudis Catalans. p. 51. ISBN 847283610X. 
  171. ^ Brown, Paul L.; Mompean, Federico J.; Perrone, Jane; Illemassène, Myriam (2005). Chemical Thermodynamics of Zirconium. Gulf Professional Publishing. p. 144. ISBN 0-444-51803-7. http://books.google.com/?id=DvqwTdVhjMEC&pg=PA144. 
  172. ^ Yaws & Braker 2001, p. 397.
  173. ^ Bartlett, Neil; Lohmann, D. H. (1962). "Dioxygenyl hexafluoroplatinate (V), O2+[PtF6]". Proceedings of the Chemical Society (Chemical Society) (3): 115. doi:10.1039/PS9620000097. 
  174. ^ Bartlett, Neil (1962). "Xenon hexafluoroplatinate (V) Xe+[PtF6]". Proceedings of the Chemical Society (Chemical Society) (6): 218. doi:10.1039/PS9620000197. 
  175. ^ Vogt, T.; Fitch, A. N.; Cockcroft, J. K. (1994). "Crystal and molecular structures of rhenium heptafluoride". Science 263 (5151): 1265–67. Bibcode 1994Sci...263.1265V. doi:10.1126/science.263.5151.1265. PMID 17817431. 
  176. ^ a b c d Bayerische Julius-Maximilians-Universität Würzburg (2006). The Highest Oxidation States of the 5d Transition Metals: a Quantum-Chemical Study (Report). Chemical Society. pp. 1–140. 
  177. ^ a b Lemal, D. M. (2004). "Perspective on fluorocarbon chemistry". Journal of Organic Chemistry 69 (1): 1–11. doi:10.1021/jo0302556. PMID 14703372. 
  178. ^ (German)Otto Ruff, Joseph Fischer, Fritz Luft (1928). "Das Stickstoff-3-fluorid". Zeitschrift für anorganische und allgemeine Chemie 172 (1): 417–25. doi:10.1002/zaac.19281720132. 
  179. ^ Yaws & Braker 2001, p. 651.
  180. ^ Lewars 2008, p. 57.
  181. ^ Musher, J. I. (1969). "The chemistry of hypervalent molecules". Angewandte Chemie International Edition 8: 54–68. doi:10.1002/anie.196900541. 
  182. ^ Brinck, T.; Murray, J. S.; Politzer, P. (1993). "A computational analysis of the bonding in boron trifluoride and boron trichloride and their complexes with ammonia". Inorganic Chemistry 32 (12): 2622–25. doi:10.1021/ic00064a008. 
  183. ^ Raghavan, P.S. (1997). Concepts and Problems in Physical Chemistry. p. 422. ISBN 8171413935. 
  184. ^ (German)Hebecker, C. (1971). "Zur Kristallstruktur von Wismutpentafluorid" (in German). Zeitschrift für anorganische und allgemeine Chemie 384 (2): 111–14. doi:10.1002/zaac.19713840204. 
  185. ^ Wells, A. F. (1984). Structural Inorganic Chemistry (Fifth ed.). ISBN 0198553706. 
  186. ^ Christe, K. O.; Wilson, W. W. (1986). "Synthesis and characterization of tetrafluoroammonium(1+) tetrafluorobromate(1-) and tetrafluoroammonium(1+) tetrafluorooxobromate(1-)". Inorganic Chemistry 25 (11): 1904–06. doi:10.1021/ic00231a038. 
  187. ^ Smith, D. F. (1963). "Chlorine pentafluoride". Science 141 (3585): 1039–40. Bibcode 1963Sci...141.1039S. doi:10.1126/science.141.3585.1039. PMID 17739492. 
  188. ^ Hyde, G. A.; Boudakian, M. M. (1968). "Synthesis routes to chlorine and bromine pentafluorides". Inorganic Chemistry 7 (12): 2648–49. doi:10.1021/ic50070a039. 
  189. ^ Schumb, W. C.; Lynch, M. A., Jr. (1950). "Iodine Heptafluoride". Industrial & Engineering Chemistry 42 (7): 1383–86. doi:10.1021/ie50487a035. 
  190. ^ Zuckerman, J. J.; Hagen, A. P. (1989). Inorganic Reactions and Methods, the Formation of Bonds to Halogens. John Wiley and Sons. p. 31. ISBN 0471186562. 
  191. ^ Wiberg, Wiberg & Holleman 2001, pp. 392–393.
  192. ^ Chemical and Engineering News as cited by Michael Barnes. "Neil Bartlett, emeritus professor of chemistry, dies at 75". University of California Newsroom. http://www.universityofcalifornia.edu/news/article/18376. Retrieved 2011-12-24. 
  193. ^ Lewars 2008, pp. 70–71.
  194. ^ Chaban, Galina M.; Lundell, Jan; Gerber, R. Benny (2001). "Lifetime and decomposition pathways of a chemically bound helium compound". The Journal of Chemical Physics 115 (16): 7341–44. Bibcode 2001JChPh.115.7341C. doi:10.1063/1.1412467. 
  195. ^ Lewars 2008, pp. 70–78.
  196. ^ Steigerwald, F.; Walter, W.; Langhoff, H.; Hammer, W. (1988). "Emission spectrum and formation kinetics of neon fluoride". Zeitschrift für Physik D 7 (4): 379–382. Bibcode 1988ZPhyD...7..379S. doi:10.1007/BF01439807. 
  197. ^ Stump, B.; Eberle, C.; Schweizer, W. B.; Kaiser, M.; Brun, R.; Krauth-Siegel, R. L.; Lentz, D.; Diederich, F. (2009). "Pentafluorosulfanyl as a novel building block for enzyme inhibitors: trypanothione reductase inhibition and antiprotozoal activities of diarylamines". ChemBioChem 10 (1): 79. doi:10.1002/cbic.200800565. PMID 19058274. 
  198. ^ Basting, D.; Marowsky, G. (2005). "Historical Review of the Excimer Laser Development". Excimer Laser Technology. Springer. pp. 8–20. ISBN 9783540200567. 
  199. ^ a b Wiberg, Wiberg & Holleman 2001, p. 438.
  200. ^ a b (German)Riedel, Erwin; Janiak, Christoph (2007). Anorganische Chemie (7th ed.). Walter de Gruyter. p. 393. ISBN 3110189038. 
  201. ^ a b c d Lehmann, J. F.; Dixon, D. A.; Schrobilgen, G. J. (2001). "X-ray Crystal Structures of α-KrF2, [KrF][MF6] (M = As, Sb, Bi), [Kr2F3][SbF6]·KrF2, [Kr2F3]2[SbF6]2·KrF2, and [Kr2F3][AsF6]·[KrF][AsF6]; Synthesis and Characterization of [Kr2F3][PF6nKrF2; and Theoretical Studies of KrF2, KrF+, Kr2F+
    3
    , and the [KrF][MF6] (M = P, As, Sb, Bi) Ion Pairs". Inorganic Chemistry (American Chemical Society) 40 (13): 3002–17. doi:10.1021/ic001167w. PMID 11399167. 
  202. ^ Grosse, A. V.; Kirshenbaum, A. D.; Streng, A. G.; Streng, L. V. (1963). "Krypton Tetrafluoride: Preparation and Some Properties". Science 139 (3559): 1047–1048. Bibcode 1963Sci...139.1047G. doi:10.1126/science.139.3559.1047. PMID 17812982.  edit
  203. ^ Prusakov, V. N.; Sokolov, V. B. (1971). "Krypton difluoride". Soviet Atomic Energy 31 (3): 990–999. doi:10.1007/BF01375764.  edit
  204. ^ a b Dixon, D. A.; Wang, T. H.; Grant, D. J.; Peterson, K. A.; Christe, K. O.; Schrobilgen, G. J. (2007). "Heats of Formation of Krypton Fluorides and Stability Predictions for KrF4and KrF6from High Level Electronic Structure Calculations". Inorganic Chemistry 46 (23): 10016–10021. doi:10.1021/ic701313h. PMID 17941630.  edit
  205. ^ Mackay, Kenneth Malcolm; Mackay, Rosemary Ann; Henderson, W. (2002). Introduction to modern inorganic chemistry (6th ed.). CRC Press. pp. 497–501. ISBN 0748764208. http://books.google.com/?id=LpJPWKT3PNcC&pg=PA497. Retrieved 2011-05-03. 
  206. ^ Christe, K. O. et al. (1995). "On the Structure of the [XeOF5] Anion and of Heptacoordinated Complex Fluorides Containing One or Two Highly Repulsive Ligands or Sterically Active Free Valence Electron Pairs". Inorganic Chemistry 34 (7): 1868–74. doi:10.1021/ic00111a039. 
  207. ^ Sharma, R. K. (1970). Chemistry of hydrides and carbides. Discovery Publishing House. p. 229. ISBN 8183562272. 
  208. ^ Pitzer, Kenneth S. (1975). "Fluorides of radon and element 118". Journal of the Chemical Society, Chemical Communications (18): 760b–61. doi:10.1039/C3975000760b. 
  209. ^ Han, Young-Kyu; Lee, Yoon Sup (1999). "Structures of RgFn (Rg = Xe, Rn, and element 118. n = 2, 4.) Calculated by two-component spin-orbit methods. A spin-orbit induced isomer of (118)F4". Journal of Physical Chemistry A 103 (8): 1104–08. doi:10.1021/jp983665k. 
  210. ^ Barber, Robert C.; Karol, Paul J.; Nakahara, Hiromichi; Vardaci, Emanuele; Vogt, Erich W. (2011). "Discovery of the elements with atomic numbers greater than or equal to 113 (IUPAC Technical Report)". Pure and Applied Chemistry: 1. doi:10.1351/PAC-REP-10-05-01. 
  211. ^ Wang, Xuefang; Andrews, Lester; Riedel, Sebastian; Kaupp, Martin (2007). "Mercury is a transition metal: The first experimental evidence for HgF4". Angewandte Chemie 119 (44): 8523–27. doi:10.1002/ange.200703710. 
  212. ^ Jensen, William B. (2008). "Is mercury now a transition element?". Journal of Chemical Education 85 (9): 1182–1183. Bibcode 2008JChEd..85.1182J. doi:10.1021/ed085p1182. http://jchemed.chem.wisc.edu/Journal/Issues/2008/Sep/abs1182.html. 
  213. ^ Emeléus, H. J.; Sharpe, A. G. (1983). Advances in Inorganic Chemistry and Radiochemistry. Academic Press. p. 83. ISBN 0120236273. 
  214. ^ (Russian) Timakov, A. A.; Prusakov, V. N.; Drobyshevskii, Y. V. (1986). Doklad akademii nauk SSSR. USSR Academy of Sciences. p. 125. 
  215. ^ Riedel, S.; Kaupp, M. (2006). "Has AuF7 been made?". Inorganic Chemistry 45 (3): 1228–34. doi:10.1021/ic051944y. PMID 16441134. 
  216. ^ Aullón, G.; Alvarez, S. (2007). "On the existence of molecular palladium(VI) compounds: Palladium hexafluoride". Inorganic Chemistry 46 (7): 2700–03. doi:10.1021/ic0623819. PMID 17326630. 
  217. ^ Weinstock, B.; Claassen, H. H.; Malm, J. G. (1957). "Platinum hexafluoride". Journal of the American Chemical Society 79 (21): 5822–32. doi:10.1021/ja01578a073. 
  218. ^ Popova, T. V.; Aksenova, N. V. (2003). "Complexes of copper in unstable oxidation states". Russian Journal of Coordinated Chemistry 29 (11): 743–65. doi:10.1023/B:RUCO.0000003432.39025.cc. 
  219. ^ (Russian)Popov, N. V.; Kiselev, Y. M.; Sukhoverkhov, V. F.; Spitsyn, V. I. (1987). Doklad akademii nauk SSSR. USSR Academy of Sciences. p. 628. 
  220. ^ Taylor, J. C.; Wilson, P. W. (1973). "The structures of fluorides—IV : A neutron diffraction study of K2NiF6". Journal of Inorganic and Nuclear Chemistry 36 (7): 1561–63. doi:10.1016/0022-1902(74)80623-8. 
  221. ^ Drews, T; Supeł, J; Hagenbach, A; Seppelt, K. (May 2006). "Solid state molecular structures of transition metal hexafluorides". Inorganic Chemistry 45 (9): 3782–88. doi:10.1021/ic052029f. PMID 16634614. 
  222. ^ Haire, Richard G. (2006). "Transactinides and the future elements". In Morss; Edelstein, Norman M.; Fuger, Jean. The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Springer Science+Business Media. p. 1723. ISBN 1-4020-3555-1. 
  223. ^ Singh, B.; Srivastava, S. (1991). "Kinetics and mechanism of ruthenium tetroxide catalysed oxidation of cyclic alcohols by bromate in a base". Transition Metal Chemistry 16: 466–68. doi:10.1007/BF01129466. 
  224. ^ O'Hagan, D. (2008). "Understanding organofluorine chemistry. An introduction to the C–F bond". Chemical Society Reviews 37 (2): 308–19. doi:10.1039/b711844a. PMID 18197347. 
  225. ^ a b Sukornick, B. (1989). "Potentially acceptable substitutes for the chlorofluorocarbons". International Journal of Thermophysics 10 (3): 553–61. Bibcode 1989IJT....10..553S. doi:10.1007/BF00507978. 
  226. ^ McMurry, John (2009). Organic Chemistry: With Biological Applications. Mary Finch. p. 618. ISBN 9780495391456. http://books.google.com/?id=h-CVwu_tk7gC&pg=PA618. Retrieved 2011-04-09. 
  227. ^ University of California, Berkeley. "Research into Gecko Adhesion". Archived from the original on 2007-10-14. http://web.archive.org/web/20071014063923/http://socrates.berkeley.edu/~peattiea/research_main.html. Retrieved 2011-04-29. 
  228. ^ Autumn, Kellar (2006). "Properties, principles, and parameters of the gecko adhesive system". Biomedical and Life Sciences: 225–56. doi:10.1007/978-3-540-31049-5_12. 

Bibliography

  • Aigueperse, Jean; Mollard, Paul; Devilliers, Didier; Chemla, Marius; Faron, Robert; Romano, Renée; Cuer, Jean Pierre (2005). Ullmann. ed. Encyclopedia of Industrial Chemistry. Wiley-VCH. p. 35. doi:10.1002/14356007. ISBN 9783527306732. 
  • Dean, John A. (1999). Lange's handbook of chemistry (Fifteenth edition). McGraw-Hill, Inc. ISBN 0070161909. 
  • Greenwood, N. N.; Earnshaw, A. (1998). Chemistry of the Elements (second edition). Butterworth Heinemann. ISBN 0750633654. 
  • Lewars, Errol G. (2008). Modelling Marvels. Springer. ISBN 1402069723. http://books.google.com/?id=IoFzgBSSCwEC&pg=PA70. Retrieved 2011-05-07. 
  • Lide, David R. (1992). Handbook of Chemistry and Physics (73rd edition, 1992–1993, special student edition). CRC Press. ISBN 0849305667. 
  • (Russian) Lidin, P. A.; Molochko, V. A.; Andreeva, L. L. (2000). Knimicheskiye svoystva neorganicheskikh veshchestv (Chemical properties of inorganic substances). Khimiya. ISBN 5-7245-1163-0. 
  • Wiberg, Egon; Wiberg, Nils; Holleman, Arnold Frederick (2001). Inorganic chemistry. Academic Press. ISBN 9780123526519. http://books.google.com/books?id=Mtth5g59dEIC. Retrieved 2011-03-03. 
  • Yaws, Carl L.; Braker, William (2001). "Fluorine". Matheson Gas Data Book, Book 2001. McGraw-Hill Professional. ISBN 9780071358545.