Jump to content

Fluorine

This is a good article. Click here for more information.
From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by R8R (talk | contribs) at 16:44, 28 December 2012 (→‎Highest oxidation states: fluorine versus oxygen). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Fluorine, 9F
Small sample of pale yellow liquid fluorine condensed in liquid nitrogen
Liquid fluorine (F2 at extremely low temperature)
Fluorine
Pronunciation
Allotropesalpha, beta (see Allotropes of fluorine)
Appearancegas: very pale yellow
liquid: bright yellow
solid: alpha is opaque, beta is transparent
Standard atomic weight Ar°(F)
Fluorine in the periodic table
Hydrogen Helium
Lithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine Neon
Sodium Magnesium Aluminium Silicon Phosphorus Sulfur Chlorine Argon
Potassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine Xenon
Caesium Barium Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury (element) Thallium Lead Bismuth Polonium Astatine Radon
Francium Radium Actinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium Rutherfordium Dubnium Seaborgium Bohrium Hassium Meitnerium Darmstadtium Roentgenium Copernicium Nihonium Flerovium Moscovium Livermorium Tennessine Oganesson


F

Cl
oxygenfluorineneon
Atomic number (Z)9
Groupgroup 17 (halogens)
Periodperiod 2
Block  p-block
Electron configuration[He] 2s2 2p5[3]
Electrons per shell2, 7
Physical properties
Phase at STPgas
Melting point(F2) 53.48 K ​(−219.67 °C, ​−363.41 °F)[4]
Boiling point(F2) 85.03 K ​(−188.11 °C, ​−306.60 °F)[4]
Density (at STP)1.696 g/L[5]
when liquid (at b.p.)1.505 g/cm3[6]
Triple point53.48 K, ​.252 kPa[7]
Critical point144.41 K, 5.1724 MPa[4]
Heat of vaporization6.51 kJ/mol[5]
Molar heat capacityCp: 31 J/(mol·K)[6] (at 21.1 °C)
Cv: 23 J/(mol·K)[6] (at 21.1 °C)
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, 0[8] (oxidizes oxygen)
ElectronegativityPauling scale: 3.98[3]
Ionization energies
  • 1st: 1681 kJ/mol
  • 2nd: 3374 kJ/mol
  • 3rd: 6147 kJ/mol
  • (more)[9]
Covalent radius64 pm[10]
Van der Waals radius135 pm[11]
Color lines in a spectral range
Spectral lines of fluorine
Other properties
Natural occurrenceprimordial
Crystal structurecubic
Cubic crystal structure for fluorine
Thermal conductivity0.02591 W/(m⋅K)[12]
Magnetic orderingdiamagnetic (−1.2×10−4)[13][14]
CAS Number7782-41-4[3]
History
Namingafter the mineral fluorite, itself named after Latin fluo (to flow, in smelting)
DiscoveryAndré-Marie Ampère (1810)
First isolationHenri Moissan[3] (June 26, 1886)
Named by
Isotopes of fluorine
Main isotopes Decay
abun­dance half-life (t1/2) mode pro­duct
18F trace 109.734 min β+ 18O
19F 100% stable
 Category: Fluorine
| references

Fluorine (symbol F) is the chemical element with atomic number 9. It is the lightest halogen. At standard pressure and temperature, fluorine is a pale yellow gas composed of diatomic molecules, F
2
. Fluorine is the most electronegative element and is extremely reactive, requiring great care in handling. It has a single stable isotope, fluorine-19.

In stars fluorine is rare compared to other light elements. In Earth's crust fluorine is the thirteenth most abundant element. Fluorine's most important mineral, fluorite, was first formally described in 1530, in the context of smelting. The mineral's name derives from the Latin verb fluo, which means "flow", because fluorite was added to metal ores to lower their melting points. Suggested as a chemical element in 1811, fluorine was named after the source mineral, but resisted many attempts to isolate the element. In 1886, French chemist Henri Moissan succeeded. His method of electrolysis remains the industrial production method for fluorine gas. The main use of elemental fluorine, uranium enrichment, was developed during the Manhattan Project.

Because of the difficulty in making elemental fluorine, most fluorine used in commerce is never converted to free fluorine. Instead, hydrofluoric acid is the key intermediate for the $16 billion per year global fluorochemical industry. The fluorides of low charged metals are ionic compounds (salts); those of high charged metals are volatile molecular compounds. The largest uses of inorganic fluorides are steel making and aluminium refining.

Organic fluorine compounds tend to have high chemical and thermal stability. The largest commercial use is in refrigerant gases (the many types of "Freons"). Although traditional chlorofluorocarbons are widely banned, the replacement gases still contain fluorine. (It is chlorine that depletes ozone, not fluorine.) Polytetrafluoroethylene (Teflon) is the most important fluoropolymer and is used in electrical insulation, chemical-resistant parts, stadium roofs, and cookware. A growing fraction of modern pharmaceuticals contain fluorine; Lipitor and Prozac are prominent examples. While a few plants and bacteria synthesize organofluorine poisons, fluorine has no metabolic role in mammals. The fluoride ion, when directly applied to teeth, reduces decay and for this reason is used in toothpaste and municipal water fluoridation.

Characteristics

Physical properties

Fluorine forms diatomic molecules that are gaseous at room temperature. The density is about 1.3 times that of air.[16][note 1] This molecule has a relatively small enthalpy of dissociation, about two-thirds of the dichlorine value.[18] This can be explained in different ways (see below).

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.[19] The element has a "pungent" characteristic odor that is noticeable in concentrations as low as 20 ppb.[20]

An 1892 observation of fluorine gas (middle), compared to air (left) and chlorine (right): the observation was done looking down 5 meter long tubes.
The color of fluorine gas (2), compared to air (1) and chlorine (3). The observation was done in 1892, looking down 5 meter tubes.

Fluorine condenses to a bright yellow liquid at −188 °C (−307 °F),[21] close to the condensation temperatures of oxygen and nitrogen. Fluorine solidifies at −220 °C (−363 °F)[21] into a cubic structure, called beta-fluorine. This phase is transparent and soft, with significant disorder of the molecules. At −228 °C (−378 °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's than to the other halogens'.[22][23]

angled lines showing linear pressure temp relations of the lower phase boundaries A parallelogram-shaped outline with space-filling diatomic molecules (joined circles) arranged in two layers
Phase diagram of fluorine Alpha-fluorine crystal structure

Electronic structure

A fluorine atom has nine protons and nine electrons, one fewer than neon, arranged in the electronic configuration [He]2s22p5.[24] 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 high effective nuclear charge. Fluorine has a relatively small covalent radius, slightly less than 60 picometers, which is less than its neighbors, oxygen and neon.[25]

A bunch of rings showing valence and non-valence electron shells
Structure of the fluorine atom

Fluorine is reluctant to ionize and instead has an attraction for one more electron to achieve the extremely stable neon-like arrangement.[24] 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.[26] 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.[27]

Although an individual fluorine atom has one unpaired electron, in molecular fluorine all the electrons are paired. Because of this, elemental fluorine is diamagnetic (slightly repelled by magnets). In contrast, the molecules of neighbor element oxygen, with two unpaired electrons per molecule, are paramagnetic (attracted to magnets).[13] The measured value of fluorine's magnetic susceptibility is −9.6×10−6 (cgs), which is close to theoretical predictions. The experimental result was not accomplished until 1999 because of the difficulties in handling fluorine gas as well as the need to specially purify the fluorine of any trace of paramagnetic oxygen.[28]

As mentioned, the molecule is relatively easily broken up. The fluorine molecule can be said to be unique among the halogens in having a bond order of exactly 1 (exactly single bonded diatomic molecules). For comparison, chlorine has a bond order of 1.12.[29] In the heavier halogens, the presence of a d-subshell (first occurring at the third energy sublevel) allows for the slightly more than single bonding. Fluorine's outermost electrons are located at the second energy sublevel where there is no d subshell.[18] Another approach would be size-caused increased electron and internuclear repulsions, as well as decreased bonding orbital overlap.[18]

Chemical reactivity

Fluorine's chemistry is dominated by its tendency to gain an electron. It is the most electronegative element and a strong oxidant. 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.[30]

Halogen bond energies (kJ/mol)[31]
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 gas is highly reactive with other substances both because of the strong bonds it forms with other atoms and because of the relative weakness of the fluorine–fluorine bond. That bond energy is significantly weaker than those of dichlorine or dibromine molecules and similar to the easily cleaved oxygen–oxygen bonds of peroxides or nitrogen–nitrogen bonds of hydrazines.[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.[20]

Reactions with fluorine are often sudden or explosive. Many generally non-reactive substances such as powdered steel, glass fragments and asbestos fibers are readily consumed by cold fluorine gas. Wood and even water burn with flames when subjected to a jet of fluorine, without the need for a spark.[16][33]

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

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). The alkali metals react with fluorine with a bang (small explosion), while the alkaline earth metals react at room temperature as well but not as aggressively. 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]

Isotopes

Fluorine occurs naturally on Earth exclusively in the form of its only stable isotope, fluorine-19,[44] which makes the element monoisotopic and mononuclidic. Seventeen radioisotopes have been synthesized: mass numbers 14–18 and 20–31.[45] 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.[46]

The lightest fluorine isotopes, those with mass numbers of 14–16, decay via electron capture. 17F and 18F 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.[45]

Only one nuclear isomer (long-lived excited nuclear state), fluorine-18m, has been characterized.[47] 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.[47]

Origin and occurrence

In the universe

Abundance in the Solar System[48]
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.[49][50]

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.[49][50]

In space, fluorine commonly combines with hydrogen to form hydrogen fluoride. (This compound has been suggested as a proxy to enable tracking reservoirs of hydrogen in the universe.)[51] In addition to HF, monatomic fluorine has been observed in the interstellar medium.[52][53] Fluorine cations have been seen in planetary nebulae and in stars, including our Sun.[54]

On Earth

Fluorine is the thirteenth most common element in Earth's crust, comprising between 600 and 700 ppm of the crust by mass. Because of its reactivity, it is essentially only found in compounds. Three minerals exist that are industrially relevant sources: fluorite, fluorapatite, and cryolite.[55][56]

  • Fluorite (CaF2), also called fluorspar or Blue John, is the main source of commercial fluorine. Fluorite is a colorful mineral associated with hydrothermal deposits. It is common and found worldwide. China supplies more than half of the world's demand; Mexico is the second-largest producer. The United States produced most of the world's fluorite in the early 20th century, but the last mine, in Illinois, shut down in 1995.[56][57][58][59]
  • Fluorapatite (Ca5(PO4)3F) is mined along with other apatites for its phosphate content and is used mostly for production of fertilizers. Most of the Earth's fluorine is bound in this mineral, but because the percentage within the mineral is low (3.5%), the fluorine is discarded as waste. Only in the United States is there significant recovery. There the hexafluorosilicates produced as byproducts are used to supply municipal water fluoridation.[56]
  • Cryolite (Na3AlF6) is the least abundant of the three, but is a concentrated source of fluorine. It was formerly used directly in aluminium production. However, the main commercial mine, on the west coast of Greenland, closed in 1987.[56]
Major fluorine-containing minerals
pink globular mass with crystal facets Long prism-like crystal, without luster, at an angle coming out of aggregate-like rock A parallelogram-shaped outline with space-filling diatomic molecules (joined circles) arranged in two layers
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.[56]

Organofluorines have been observed in volcanic eruptions and in geothermal springs. Their ultimate origin is said to vary from physical formation under geological conditions to initial biological production and deposition in sediments. However, the provenance is still being studied, as is the natural organofluorine distribution. They are not found in large quantites (compare also the number of known natural organofluorines, 30, to that of organochlorines, 2150), so they are not commercially important source of fluorine.[60]

The possibility of small amounts of gaseous fluorine within crystals has been debated for many years. One form of fluorite, antozonite, has a smell suggestive of fluorine when crushed. The mineral also has a dark black color, perhaps from free calcium (not bonded to fluoride). In 2012, a study reported detection of trace quantities (0.04% by weight) of diatomic fluorine in antozonite. It was suggested that radiation from small amounts of uranium within the crystals had caused the free fluorine defects.[61]

History

man with long stick-ladel thingie, turning towards a domed, hot, brick oven thingie with sluices coming out of it
Smelting illustration from Agricola's De re metallica, where fluorite was first described

The word "fluorine" derives from the Latin stem of the main source mineral, fluorite, which was first mentioned in 1529 by Georgius Agricola, who described it as a flux—an additive that helps melt ores and slags during smelting.[62][63] Fluorite stones were called schone flusse in the German of the time. Agricola, writing in Latin but describing 16th century industry, invented several hundred new Latin terms. For the schone flusse stones, he used the Latin noun fluores, "fluxes", because they made metal ores flow when in a fire. After Agricola, the name for the mineral evolved to fluorspar (still commonly used) and then to fluorite.[59][64][65]

Some sources claim that the first production of hydrofluoric acid was by Heinrich Schwanhard, a German glass cutter, in 1670.[66] A peer-reviewed study of Schwanhard's writings, though, showed no specific mention of fluorite and only discussion of an extremely strong acid. It was hypothesized that this was probably nitric acid or aqua regia, both capable of etching soft glass.[67][68] 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.[69] In 1771, Swedish chemist Carl Wilhelm Scheele repeated this reaction.[69] In 1810, French physicist André-Marie Ampère suggested that the acid was a compound of hydrogen with an unknown element, analogous to chlorine.[70] Fluorite was then shown to be mostly composed of calcium fluoride.[71]

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. (Greek, Russian, and several other languages use the name ftor or derivatives, which was suggested by Ampère and comes from the Greek φθόριος (phthorios), meaning "destructive".)[72] The new Latin name (fluorum) gave the element its current symbol, F, although the symbol Fl is seen in early papers.[73] The symbol Fl is now used for the super-heavy element flerovium.[74]

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 its electrolytical preparation was hard to perform and because the gas reacted with most materials. The generation of elemental fluorine proved to be exceptionally dangerous, killing or blinding several early experimenters. Jean Dussaud referred to these men as "fluorine martyrs", a term still used.[71]

Edmond Frémy thought that passing electric current through pure hydrofluoric acid might work. Previously, hydrogen fluoride was only available in a water solution. Frémy therefore devised a method for producing dry hydrogen fluoride by acidifying potassium bifluoride (KHF2). Unfortunately, pure hydrogen fluoride did not pass an electric current.[66]

painting of bearded man in a lab with an electroylsis rig
1891 graphic showing Moissan and the actual apparatus used to isolate fluorine

French chemist Henri Moissan, formerly one of Frémy's students, continued the search. After trying many different approaches, he built on Frémy's earlier attempt by combining potassium bifluoride and hydrogen fluoride. The resultant solution conducted electricity. Moissan also constructed especially corrosion-resistant equipment: containers crafted from a mixture of platinum and iridium (more chemically resistant than pure platinum) with fluorite stoppers. After 74 years of effort by many chemists, on 26 June 1886, Moissan reported the isolation of elemental fluorine.[66][75] Moissan's report to the French Academy of making fluorine showed appreciation for the feat:

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[76]

Moissan later devised a less expensive apparatus for making fluorine: copper equipment coated with copper fluoride. In 1906, two months before his death, Moissan received the Nobel Prize in chemistry for his fluorine isolation as well as the invention of the electric arc furnace.[77][78][79]

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 formed a joint venture in 1930 to market the new product; in 1949 DuPont took over the business. Freon proved to be a marketplace hit, rapidly replacing earlier, more toxic, refrigerants and growing the overall market for kitchen refrigerators.[69][80][81][82]

In 1938, polytetrafluoroethylene (DuPont brand name Teflon) was discovered by accident by a recently-hired DuPont PhD, Roy J. Plunkett. While working with tetrafluoroethylene gas, he noticed missing weight. Scraping down his container, he found white flakes of a polymer new to the world. 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.[69][81][83]

top level view of large industrial buildings
The Manhattan Project's K-25 gaseous diffusion plant in Oak Ridge, Tennessee

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.[84] 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.[67] Because uranium hexafluoride releases small quantities of corrosive fluorine, the separation plants were built with special materials. All pipes were coated with nickel; joints and flexible parts were fabricated from Teflon.[81]

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.[85] In 1969, Robert Gore developed an expanded PTFE membrane which led to the large Gore-tex business in breathable rainwear. The company developed many other uses of PTFE.[86]

In the 1970s and 1980s, 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.[87][88]

Industry and applications

The global market for fluorochemicals was about $16 billion per year as of 2006.[89] Historically, the industry has grown a few percent per year and is predicted to do so in the future. Although fluorochemical demand contracted during the 2008–2009 global recession, the industry was predicted to reach 2.6 million tons per year by 2015.[90]

The largest market is the United States. Western Europe is the second largest. Asia Pacific is the fastest growing region of production.[90] China in particular has experienced significant growth as a fluorochemical market and is becoming a producer of them as well.[91][92]

Fluorite mining (the main source of fluorine) was estimated in 2003 to be a $550 million industry, extracting 4.5 million tons per year. Most ores must be processed to concentrate the fluorite from other minerals by various methods of flotation separation. However, only about 1% of mined fluorite is ever converted to elemental fluorine.[69]

FluoriteFluorapatiteHydrogen fluorideMetal smeltingGlass productionFluorocarbonsSodium hexafluoroaluminatePickling (metal)Fluorosilicic acidAlkane crackingHydrofluorocarbonHydrochlorofluorocarbonsChlorofluorocarbonTeflonWater fluoridationUranium enrichmentSulfur hexafluorideTungsten hexafluoridePhosphogypsum
Fluorine industry supply chain: major sources, intermediates and applications. Click for links to related articles.

Inorganic fluorides

Mined fluorite is separated into three main grades, with about equal production of each. Acidspar is at least 97% CaF2; metspar is much lower purity, 60–85%; the ceramic grade is in between, 85–95%.[69][92]

Metspar is used almost exclusively for iron smelting. Acidspar is primarily converted to hydrofluoric acid (by reaction with sulfuric acid). The resultant HF is mostly used to produce organofluorides and synthetic cryolite.

About 3 kg (6.5 lb) 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 content has a tangential benefit in removing sulfur and phosphorus, but other additives such as lime are still needed. Metspar is similarly used in cast iron production and for other iron-containing alloys.[93][94]

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.[93] 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.[93]

Aluminium smelting process: cryolite (a fluoride) is required to dissolve alumnium oxide.

One third of HF (one sixth of mined fluorine) is used to make synthetic cryolite (sodium hexafluoroaluminate) and aluminium trifluoride. These compounds are used in the electrolysis of aluminium. About 23 kg (51 lb) are required for every tonne of aluminium. These compounds are also used as a flux for glass.[93]

Fluorosilicates are the next most significant inorganic fluorides formed from HF. The most common one, that of sodium, is used for water fluoridation, as an intermediate for synthetic cryolite and silicon tetrafluoride, and for treatment of effluents in laundries. Other fluorosilicates have their specific uses.[95]

MgF2 and, to a lesser extent, other alkaline earth difluorides are specialty optical materials.[96] Magnesium difluoride is widely used as an antireflection coating for spectacles and optical equipment.[97] The compound is also a component in newly devised constructions (negative index metamaterials) which are the subject of "invisibility" research. The layered structures can curve light around objects.[98][99][100]

Other inorganic fluorides made in large quantities include cobalt difluoride (for organofluorine synthesis), nickel difluoride (electronics), lithium fluoride (a flux), sodium fluoride (water fluoridation), potassium fluoride (flux), and ammonium fluoride (various).[93] Sodium and potassium bifluorides are significant to the chemical industry.[101]

Fluorocarbons

Making organic fluorides is the main use for hydrofluoric acid, consuming over 40% of it (over 20% of all mined fluorite). Within organofluorides, refrigerant gases are still the dominant segment, consuming about 80% of HF. Even though chlorofluorocarbons are widely banned, the replacement refrigerants are often other fluorinated molecules. Fluoropolymers are less than one quarter the size of refrigerant gases in terms of fluorine usage, but are growing faster.[90][93] Fluorosurfactants are a small segment in mass but are significant economically because of very high prices.

Refrigerant gases

Traditionally chlorofluorocarbons (CFCs) were the predominant fluorinated organic chemical. CFCs are identified by a system of numbering that explains the amount of fluorine, chlorine, carbon and hydrogen in the molecules. The term Freon has been colloquially used for CFCs and similar halogenated molecules, though strictly speaking this is just a DuPont brand name, and 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).[93]

a sailer in coveralls wipes down red gas bottles
A Halon fire suppression system in a ship's machinery room

Production of CFCs grew strongly through the 1980s, primarily for refrigeration and air conditioning but also for propellants and solvents. 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-1980s peak, with remaining use primarily as an intermediate for other chemicals. The banning of CFCs initially depressed the overall demand for fluorite but 21st century production of the source mineral has recovered to 1980s levels.[93]

Hydrochlorofluorocarbons (HCFCs) and hydrofluorocarbons (HFCs) now serve as replacements for CFC refrigerants; 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).[93]

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.[102]

Fluoropolymers

Fluoropolymers are less than 0.1% of all polymers produced in terms of weight. They are more expensive and have higher growth rates than average polymers. As of about 2006–2007, estimates of the global fluoropolymer production varied from over 100,000 to 180,000 metric tons per year. Yearly revenue estimates ranged from over $2.5 billion to over $3.5 billion.[103][104]

PTFE (polytetrafluoroethylene) is 60–80% of the world's fluoropolymer production on a weight basis.[104] The term Teflon is sometimes used generically for the substance, but is a DuPont brand name—other PTFE producers exist and DuPont sometimes uses the Teflon brand for other 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 tetrafluoroethylene (abbreviated TFE), the starting point for PTFE.[105]

The largest application for PTFE is in electrical insulation. It is an excellent dielectric and very chemically stable. It is also used extensively in the chemical process industry where corrosion resistance is needed: in coating pipes, in tubing, and gaskets. Another major use is architectural fabric (PTFE-coated fiberglass cloth used for stadium roofs and such). The major consumer application is non-stick cookware.[105]

Major PTFE applications
cutaway of a coax cable (plastic around metal around PTFE around metal core) advertisement for the "Happy Pan" with DuPont Teflon vast expanse with no columns, gray roof with sections, bright green playing field
PTFE dielectric separating core and outer metal in a specialty coaxial cable First Teflon branded frying pan, 1961 The interior of the Tokyo Dome. The roof is PTFE-coated fiberglass and air-supported.[106]

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 bag house filters for industries with corrosive exhausts.[105]

bunch of tendrils
Microscopic structure of ePTFE ("Gore-Tex")

Other fluoropolymers 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.[105][107][108]

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.[109][110][111]

Fluoroelastomers are rubber-like substances that are composed of crosslinked mixtures of fluoropolymers. Viton is a prominent example. Chemical-resistant O-rings are the primary application. Fluoroelastomers tend to be more stiff than conventional elastomers, but with superior chemical and heat resistance.[105]

Surfactants

A shiny spherical drop of water on blue cloth.
Drop of water on a fabric treated with fluorinated surfactant

Fluorinated surfactants are small organofluorine molecules, principally used in stain repellents and DWR (durable water repellent). Fluorosurfactants form a large market, over $1 billion per year as of 2006. Scotchgard is a prominent brand, with over $300 million revenue in 2000.[112] Fluorosurfactants are expensive chemicals, comparable to pharmaceutical chemicals: $200–2000 per kilogram ($90–900 per pound).[113]

Fluorosurfactants make a very small part of the overall surfactant market, most of which is hydrocarbon based and much cheaper. Some potential applications (e.g. low cost paints) are unable to use fluorosurfactants because of the price impact of compounding in even small amounts of fluorosurfactant. Use in paints was only about $100 million as of 2006.[112]

DWR is a final (very thin coating) put on fabrics that makes them lightly rain resistant, that makes water bead. First developed in the 1950s, fluorosurfactants were 90% of the DWR industry by 1990. 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).[114]

Fluorine gas

For countries with available data (free-market countries), about 17,000 metric tons of fluorine are produced per year by 11 companies, all G7-resident.[115] Fluorine is relatively inexpensive, costing about US$5–8 per kilogram ($2–4 per pound) when sold as uranium hexafluoride or sulfur hexafluoride. Because of difficulties in storage and handling, the price of pure fluorine gas is much higher.[115] Processes demanding large amounts of fluorine gas generally vertically integrate and produce the gas onsite for direct use.

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.[115] 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.[93] Up to 7,000 tons per year of fluorine gas are used for this application.[115]

minaret-like electrical things with wires running around and fat at the bottom
SF6 transformers at a Russian railway

The second largest application for fluorine gas is sulfur hexafluoride, which is used as a 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 switchgear, 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.[116]

Several compounds made from elemental fluorine serve the electronics industry. Rhenium and tungsten hexafluorides are used for chemical vapor deposition of thin metal films onto semiconductors. Tetrafluoromethane, is used for plasma etching in semiconductor manufacturing, flat panel display production, and microelectromechanical systems fabrication.[117][118][119] Nitrogen trifluoride is increasingly used for cleaning equipment at display manufacturing plants. Elemental fluorine, itself, is used sometimes for cleaning equipment.[93]

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.[93]

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.[120][121]

Production of fluorine gas

A machine room
Fluorine cell room at F2 Chemicals, Preston, England.

Commercial producers of fluorine gas continue to use the method of electrolysis pioneered by Moissan, with some modifications in the cell design. Owing to the gas's corrosiveness, special containment materials and handling precautions are required. Chemical routes to the elemental form were published in 1986.

Electrolytic synthesis

Several thousand tons of elemental fluorine are produced annually by electrolysis of potassium bifluoride in hydrogen fluoride.[93] 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).[69] 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.[122]

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.[123]

Handling

Pure fluorine gas may be stored in steel cylinders where the inside surface is passivated by a metal fluoride layer that resists further attack.[122][69] Passivated steel will withstand fluorine provided the temperature is kept below 200 °C (400 °F). Above that temperature, nickel is required. Regulator valves are made of nickel. Fluorine piping is generally made of nickel or Monel (nickel-copper alloy).[124] Care must be taken to passivate all surfaces frequently and to exclude any water or greases. In the laboratory, fluorine gas can be used in glass tubing provided the pressure is low and moisture is excluded,[124] although some sources recommend systems made of nickel, monel, and PTFE.[125]

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; however, he stated in his work that the basics were known 50 years before the actual reaction.[126] The main idea is that some metal fluoride anions do not have a neutral counterpart (or those are very unstable) and their acidifying would result in chemical oxidation, rather than formation of the expected molecules. Christe lists the following reactions as a possible way:

2 KMnO4 + 2 KF + 10 HF + 3 H2O2 → 2 K2MnF6 + 8 H2O + 3 O2
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.[127]

Biological aspects

South Africa's gifblaar is one of the few organisms that makes fluorine compounds.

Fluoride is not considered an essential mineral element for mammals and humans.[128] Small amounts of fluoride may be beneficial for bone strength, but this is an issue only in the formulation of artificial diets.[129] Food and drinking water typically contain at least small amounts of fluorides, which are naturally present.

Several other biological aspects of fluorine exist. Fluoride is widely used for prevention of dental cavities. In pharmaceuticals and agrichemicals, fluorine sees increasing use in new molecules. Both radioactive and natural fluorine isotopes are important in respective scanning applications. Oxygen-carrying perfluorocarbons present a possibility for human liquid breathing. Poisons containing fluorine are well known for killing insects and rodents and the very few organisms that incorporate fluorine in their biochemistry do so to make natural poisons.

Dental care

Fluoride ions in contact with teeth have been long thought to limit cavities by turning the forming the hydroxyapatite of teeth into less soluble fluorapatite. The more up-to-date studies show no difference in caries levels between teeth with enamel fluoridated to different degrees, while low levels of fluoride in plaque fluid and saliva definitely do help to fight the early caries. The process of fluoride absorption works only by direct contact (topical treatment). Fluoride ions that are swallowed do not benefit the teeth.[130]

white man holding plastic tray with brown goop in it and sticking a small stick into a black boy's open mouth
Topical fluoride treatment in Panama

Water fluoridation is the controlled addition of fluoride to a public water supply to reduce tooth decay.[131] Its use began in the 1940s, following studies of children in a region where water is naturally fluoridated. It is now used for about two-thirds of the U.S. population on public water systems[132] and for about 5.7% of people worldwide.[133] Although the best available evidence shows no association with adverse effects other than fluorosis (dental and, in worse cases, skeletal), most of which is mild,[134] water fluoridation has been contentious for ethical, safety, and efficacy reasons,[133] and opposition to water fluoridation exists despite its support by public health organizations.[135] The benefits of water fluoridation have lessened recently, presumably because of the availability of fluoride in other forms, but are still measurable, particularly for low income groups.[136] Systematic reviews in 2000 and 2007 showed significant reduction of cavities in children associated with water fluoridation.[137]

Sodium fluoride, tin difluoride, and, most commonly, sodium monofluorophosphate, are used in toothpaste. In 1955, the first fluoride toothpaste was introduced, in the United States. Now, almost all toothpaste in developed countries is fluoridated. For example, 95% of European toothpaste contains fluoride.[136] Gels and foams are often advised for special patient groups, particularly those undergoing radiation therapy to the head (cancer patients). The patient receives a four minute application of a high amount of fluoride. Varnishes exist that perform a similar function, but are more quickly applied. Fluoride is also contained in prescription and non-prescription mouthwashes and is a trace component of foods manufactured from fluoridated water supplies.[138]

Pharmaceuticals

large image of just a capsule with words Prozac and DISTA visible
Prozac capsules, the 20-milligram commercially available form

Several important pharmaceuticals contain fluorine.[139] 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.[20] 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 bacause of increased cell membrane penetration.[140] Although the potential of the fluorine to be released as a fluoride leaving group is dependent on its position in the molecule,[141] organofluorides are generally very stable, since the carbon–fluorine bond is strong.

Of commercialized drugs, already 20% contain fluorine,[142] as even a single atom of it improves the chemical reactivity of the drug molecule used. 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.[143] 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.[144]

Many SSRI antidepressants are fluorinated organics,[145] including citalopram, escitalopram, and fluoxetine. Fluoroquinolones are a commonly used family of broad-spectrum antibiotics.[146] 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.[147]

Fluorine-containing life science chemicals
File:Trifluralin-3D-matt-sf.png
Lipitor (atorvastatin) 5-FU (fluorouracil) Florinef (fludrocortisone) Isoflurane Trifluralin

Agrichemicals and natural poisons

An estimated 30% of agrichemical compounds contain fluorine.[148] Most of them are poisons, but a few stimulate the growth instead. The further fluorine agrichemicals usage is said to rely on improvement of synthesis reaction (lowering the costs raised by fluorine content) and level of greener chemistry consideration (fluorochemicals are more environment-friendly).[149]

Synthetic sodium fluoroacetate has been used as an insecticide but is especially effective against mammalian pests.[150] The name "1080" refers to the catalogue number of the poison, which became its brand name.[151] Fluoroacetate is similar to acetate, which has a pivotal role in the Krebs cycle (a key part of cell metabolism). Fluoroacetate halts the cycle and causes cells to be deprived of energy.[151] Several other insecticides contain sodium fluoride, which is much less toxic than fluoroacetate.[152] Currently the compound is banned.[153] Another important agrichemcial is Trifluralin. It was once very important (for example, in 1998 over a half of U.S. cotton field area was coated with the chemical[154]); however, its suspected carcinogenic properties caused some Northern European countries to ban it in 1993.[155] Currently, the whole European Union has it banned, although there was a case intended to cancel the decision.[156]

The currently used agrichemicals utilize another tactic: Instead of being poisonous themselves, they penetrate into the metabolism not to affect it directly, but to transform into actually poisonous compounds. For example, insects fed 29-fluorosigmasterol produce the fluoroacetates from it. If a fluorine is transferred to a body cell, it blocks metabolism at the position occupied.[157]

Biologically synthesized organofluorines have been found in microorganisms and plants,[60] but not in animals.[158] The most common example is fluoroacetate, with an active poison molecule identical to commercial "1080" described above. It is used as a defense against herbivores by at least 40 green plants in Australia, Brazil, and Africa;[151] other biologically synthesized organofluorines include ω-fluoro fatty acids, fluoroacetone, and 2-fluorocitrate.[158] In bacteria, the enzyme adenosyl-fluoride synthase, which makes the carbon–fluorine bond, was isolated. The discovery was touted as possibly leading to biological routes for organofluorine synthesis.[159] Aside from natural organofluorines, there is a significant runoff organofluorines content in agricultural-sites water and nearby rivers.[160]

Scanning

A rotating, transparent image of a human figure with targeted organs highlighted
Whole-body PET scan using fluorine-18

Compounds containing fluorine-18, a radioactive isotope that emits positrons, are often used in positron emission tomography (PET) scanning, because the isotope's half-life of about 110 minutes is long by positron-emitter standards. One such radiopharmaceutical is 2-deoxy-2-(18F)fluoro-D-glucose (generically referred to as fludeoxyglucose), commonly abbreviated as 18F-FDG, or simply FDG.[161] In PET imaging, FDG can be used for asessing glucose metabolism in the brain and for imaging cancer tumors. After injection into the blood, FDG is taken up by "FDG-avid" tissues with a high need for glucose, such as the brain and most types of malignant tumors.[162] Tomography, often assisted by a computer to form a PET/CT (CT stands for "computer tomography") machine, can then be used to diagnose or monitor treatment of cancers; especially Hodgkin's lymphoma, lung cancer, and breast cancer.[163]

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. Fluorine compounds typically have a fast NMR relaxation, which enables the use of fast averaging to obtain a signal-to-noise ratio similar to hydrogen-1 NMR spectra.[164] Fluorine-19 is commonly used in NMR study of metabolism, protein structures and conformational changes.[165] In addition, inert fluorinated gases have the potential to be a cheap and efficient tool for imaging lung ventilation.[166]

Oxygen transport research

Computer-generated model of nanocrystal of perflubron (red) and gentamicin (white, an antibiotic) in liquid suspension: nanocrystals represent a method of delivering water or fat soluble drugs within a perfluorochemical fluid. The particles usage is being developed to help treat babies with damaged lungs.[167]

Liquid fluorocarbons have a very high capacity for holding gas in solution. They can hold more oxygen or carbon dioxide than blood does. For that reason, they have attracted ongoing interest related to the possibility of artificial blood or of liquid breathing.[168]

Blood substitutes are the subject of research because the demand for blood transfusions grows faster than donations. In some scenarios, artificial blood may be more convenient or safe. Because fluorocarbons do not normally mix with water, they must be mixed into emulsions (small droplets of perfluorocarbon suspended in water) to be used as blood.[169][170][171] One such product, Oxycyte, has been through initial clinical trials.[172][173]

Possible medical uses of liquid breathing (which uses pure perflurocarbon liquid, not a water emulsion) involve assistance for premature babies or for burn victims (because the normal lung function is compromised). Both partial filling of the lungs and complete filling of the lungs have been considered, although only the former has any significant tests in humans. Several animal tests have been performed and some human partial liquid ventilation trials.[174] One effort, by Alliance Pharmaceuticals, reached clinical trials but was abandoned because of insufficient advantage compared to other therapies.[175]

Other posited applications include deep sea diving and space travel, applications that would both require total liquid ventilation, not partial ventilation.[176][177] The 1989 film The Abyss showed a fictional use of perfluorocarbon for human diving but also filmed a real rat surviving while cooled and immersed in perfluorocarbon.[178] Several other fictional treatments of perfluorocarbon breathing exist.

Hazards

Fluorine gas and fluoride ion

4 diagonal placards with warnings, poison, corrosive, inhalant, oxidant
U.S. hazard signs for commercially transported fluorine[179]

Elemental fluorine is highly toxic. Above a concentration of 25 ppm, fluorine causes significant irritation while attacking the eyes, respiratory tract, lungs, liver and kidneys. At a concentration of 100 ppm, human eyes and noses are seriously damaged.[180]

Soluble fluorides are moderately toxic. For sodium fluoride, the lethal dose for adults is 5–10 g, which is equivalent to 32–64 mg of elemental fluoride per kilogram of body weight.[181] The dose that may lead to adverse health effects is about one fifth the lethal dose.[182] Chronic excess fluoride consumption can lead to skeletal fluorosis, a disease of the bones that affects millions in Asia and Africa.[182][183]

The fluoride ion is readily absorbed by the stomach and intestines. Ingested fluoride forms hydrofluoric acid in the stomach. In this form, fluoride crosses cell membranes and then binds with calcium and interfere with various enzymes. Fluoride is excreted through urine. Fluoride exposure limits are based on urine testing which has determined the human body's capacity for ridding itself of fluoride.[182][184]

Historically, most cases of fluoride poisoning have been caused by accidental ingestion of insecticides containing inorganic fluoride,[185] Currently, most calls to poison control centers for possible fluoride poisoning come from the ingestion of fluoride-containing toothpaste.[182] Malfunction of water fluoridation equipment has occurred several times, including an Alaskan incident, which affected nearly 300 people and killed one.[186]

Hydrofluoric acid

left and right hands, two views, burned index fingers
HF burns, not evident until a day after

Hydrofluoric acid is a contact poison and must be handled with extreme care, far beyond that accorded to other mineral acids. 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, with 8-hour delay for 50% HF and up to 24-hour if the concentration is smaller. 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. If the burn has been initially noticed, then HF should be washed off with a forceful stream of water for ten through fifteen minutes, to prevent its further penetration into the body. Clothing used by the person burned may also exhibit danger.[187]

Once in the blood, hydrogen fluoride reacts with calcium and magnesium, resulting in electrolyte unbalance, cardiac arrhythmia, and potentially, death.[188] Formation of insoluble calcium fluoride possibly causes both a fall in calcium serum and the strong pain associated with tissue toxicity.[189] 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.[190]

Hydrofluoric acid exposure is often treated with calcium gluconate, a source of Ca2+ that binds with the fluoride ions. Skin burns can be treated with a water wash and 2.5% calcium gluconate gel[191][192] or special rinsing solutions.[193] However, because HF is absorbed, medical treatment is necessary; sometimes amputation may be required.[190]

Environmental concerns

Atmosphere

An animation showing colored representation of ozone distribution by year, above North America, through 6 steps. It starts with a lot of ozone especially over Alaska and by 2060 is almost all gone from north to south.
NASA projection of stratospheric ozone over North America if CFCs had not been banned

Chlorofluorocarbons (CFCs) and bromofluorocarbons (BFCs) have been strictly regulated via a series of international agreements, the Montreal Protocol, because they deplete the ozone layer. It is the chlorine and bromine from these molecules that cause harm, not fluorine. Because of the inherent stability of these fully halogenated molecules (which makes them so nonflammable and useful), they are able to reach the upper reaches of the atmosphere, before decomposing, and then release chlorine and bromine to attack the ozone at those altitudes.[194] Predictions are that generations will be required, even after the CFC ban, for these molecules to leave the atmosphere and for the ozone layer to recover. Early indications are that the CFC ban is working—ozone depletion has stopped and recovery is underway.[195][196]

Hydrochlorofluorocarbons (HCFCs) are current replacements for CFCs; HCFCs have about one tenth the ozone damaging potential (ODP) of CFCs.[197] They were originally scheduled for elimination by 2030 in developed nations (2040 in undeveloped). In 2007, a new treaty was signed by almost all nations to move that phaseout up by ten years because HFCs, which have no chlorine and thus zero ODP, are available.[198] Meanwhile, individual HCFCs with the highest anti-ozone depleting potential are being phased out first. For example, in 2003, HCFC-141b was phased out in the U.S. by Environmental Protection Agency regulation. Many of the other HCFCs are now being produced at a fraction of their previous production rates.[199]

Fluorocarbon gases of all sorts (CFCs, HFCs, etc.) are greenhouse gases about 4,000 to 10,000 times as potent as carbon dioxide. Sulfur hexafluoride exhibits an even stronger effect, about 20,000 times the global warming potential of carbon dioxide.[200][201]

Biopersistance

jumping dolphin
Bottlenose dolphins have high PFOS.

Because of the strength of the carbon–fluorine bond, organofluorines endure in the environment. Perfluorooctanoic acid (PFOA) and perfluorooctanesulfonic acid (PFOS), used in waterproofing sprays, are persistent global contaminants. Trace quantities of these substances have been detected worldwide, from polar bears in the Arctic to the global human population. One study indicates that PFOS levels in wildlife are starting to go down because of the recent reduced production of that chemical.[202][203]

PFOA's tissue distribution in humans is unknown, but studies in rats suggest it is likely to be present primarily in the liver, kidney, and blood. In the body, PFOA binds to a protein, serum albumin; it has been detected in breast milk and the blood of newborns. PFOA is not metabolized by the body, but is excreted by the kidneys.[202][203]

The potential health effects of PFOA are unclear. Unlike chlorinated hydrocarbons, PFOA is not lipophilic (stored in fat), nor is it genotoxic (damaging genes). While both PFOA and PFOS cause cancer in high quantities in animals, studies on exposed humans have not been able to prove an impact at current exposures. Bottlenose dolphins have some of the highest PFOS concentrations of any wildlife studied; one study suggests an impact on their immune systems.[202][203]

Compounds

Fluorine as straight (a) or bent (b) bridging ligands.[204]

Fluorine's only common oxidation state is −1.[note 2] With other atoms, fluorine forms either polar covalent bonds or ionic bonds. Most frequently, covalent bonds involving fluorine atoms are single bonds, although at least two examples of a higher order bond exist.[207] Fluoride may act as a bridging ligand between two metals in some complex molecules. Molecules containing fluorine may also exhibit hydrogen bonding.[208] Fluorine has a rich chemistry including inorganic compounds formed with hydrogen, metals, nonmetals, and even noble gases; as well as a diverse set of organic compounds.[note 3]

Inorganic

Hydrogen fluoride

HF and H2O similarities
graph showing trend-breaking water and HF boiling points: big jogs up versus a trend that is down with lower molecular weight for the other series members. graph showing humps of melting temperature, most prominent is at HF 50% mole fraction
Boiling points of the hydrogen halides (blue) and hydrogen chalcogenides (red): HF and H2O break trends. Freezing point of HF/ H2O mixtures: arrows indicate compounds in the solid state.

Fluorine combines with hydrogen to make a compound (HF) called hydrogen fluoride or, especially in the context of water solutions, hydrofluoric acid. The H-F bond type is one of the few capable of hydrogen bonding (creating extra clustering associations with similar molecules). This influences various peculiar aspects of hydrogen fluoride's properties. In some ways the substance behaves more like water, also very prone to hydrogen bonding, than one of the other hydrogen halides, such as HCl.[209][210][211]

Hydrogen bonding amongst HF molecules gives rise to high viscosity in the liquid phase and lower than expected pressure in the gas phase. Hydrogen fluoride does not boil until 20 °C in contrast to the heavier hydrogen halides which boil between −85 °C and −35 °C (−120–30 °F). HF is fully miscible with water (will dissolve in any proportion), while the other hydrogen halides have large solubility gaps with water. Hydrogen fluoride and water also form several compounds in the solid state, most notably a 1:1 compound that does not melt until −40 °C (−40 °F), which is 44 degrees Celsius (79 degrees Fahrenheit) above the melting point of pure HF.[212]

Unlike other hydrohalic acids, such as hydrochloric acid, hydrogen fluoride is only a weak acid in water solution, with acid dissociation constant (pKa) equal to 3.19.[213] HF's weakness as an aqueous acid is paradoxical considering how polar the HF bond is, much more so than the bond in HCl, HBr, or HI. The explanation for the behavior is complicated, having to do with various cluster-forming tendencies of HF, water, and fluoride ion, as well as thermodynamic issues.[note 4] At great concentrations, a property called homoconjugation is revealed. HF begins to accept fluoride ions, forming the polyatomic ions (such as bifluoride, HF
2
) and protons, thus greatly increasing the acidity of the compound.[215] Hydrofluoric acid is also the strongest of the hydrohalic acids in acetic acid and similar solvents.[216] Its hidden acidity potential is also revealed by the fact it protonates acids like hydrochloric, sulfuric, or nitric.[217] Despite its weakness, hydrofluoric acid is very corrosive, even attacking glass (hydrated only).[215]

Dry hydrogen fluoride dissolves low-valent metal fluorides readily. Several molecular fluorides also dissolve in HF. Many proteins and carbohydrates can be dissolved in dry HF and can be recovered from it. Most non-fluoride inorganic chemicals react with HF rather than dissolving.[218]

Metal fluorides

Metal fluorides have similarities with other metal halides but are more ionic. In many respects, metal fluorides differ from other metal halides (chlorides, bromides, iodides), very similar to each other. Instead, fluorides are more similar to oxides, often having similar bonding and crystal structures.[219]

The metal fluorides show broad trends based on the charge of the metal. Metals in an oxidation state of +3 or lower tend to form ionic, refractory fluorides. Metals charged +5 or higher tend to form covalently bonded fluorides, polymers or discrete molecules, and are more volatile. (The tetrafluorides are a transition zone.) The bonding variations mean that metal fluorides can be solids,[220] liquids,[221] or gases[222] at room temperature.

The solubility of fluorides varies greatly but tends to decrease as the charge on the metal ion increases. Dissolved fluorides produce basic solutions. (F- is a weak base because HF is a weak acid.)[223] Like hydroxides, fluorides may be viewed as basic, amphoteric, and acidic, with the acidity property generally increasing with the oxidation state of the metal;[29] a fluoride is a much weaker base than a hydroxide, though.[224]

Low oxidation states
checkerboard-like lattice of small blue and large yellow balls, going in three dimensions so that each ball has 6 nearest neighbors of opposite type
Sodium fluoride: sodium (purple) and fluoride (yellow) are isoelectronic, but fluoride is bigger because its nuclear charge is lower.

The alkali metals form monofluorides. All are soluble and have the sodium chloride (rock salt) structure,[225] which is also adopted by some alkaline earth oxides such as CaO.[226] Because the fluoride anion is highly basic, many alkali metal fluorides form bifluorides with the formula MHF2. They also give off the fluoride easily when reacting with an acid. Among other monofluorides, only silver(I)[73] and thallium(I)[227] fluorides are well-characterized. Both are very soluble, unlike the other halides of those metals. Another silver fluoride is from this point a "half-fluoride" (a subfluoride, compound containing less than normal saturation with fluorine), formulated as Ag
2
F
. This compound has been described as having F- and unusual "Ag½+" centers.[228]

Unlike the monofluorides, the difluorides may be either soluble or insoluble. Several transition metal difluorides, such as those of copper(II) and nickel(II), are soluble.[73] The alkaline earth metals form difluorides that are insoluble.[73] In contrast, the alkaline earth chlorides are readily soluble.[73] The insolubility of alkaline earth fluorides comes from the high negative lattice energy of the common in chemistry fluorite structure (adopted by several metal dioxides such as CeO2, UO2, ThO2, etc.), so that it is more favorable for the solid to stay solid rather than dissolve, so they are insoluble (see the table, notice the omitted minus signs). However, beryllium fluoride shows striking differences from other alkaline earth fluorides, just is beryllium different than other alkaline earth metals (tending to bond covalently, so BeF2 shows a significant covalent character). The compounds has many similarities to SiO2 (quartz): a mostly covalently bonded network solid, glasses formation (are difficult to crystallize), the same room temperature crystal structure when crystalline, etc. It is very soluble in water, unlike the other alkaline earths.[73][229][230][231][232]

Alkaline earth metals fluorides solubility-related constants[note 5]
Metal
M2+
HE
[233]
F-
HE
[234]
"MF2"
unit
HE
MF2
lattice
energies
[235]
Solubility
[236]
Be 2,455 458 3,371 3,526 soluble
Mg 1,922 458 2,838 2,978 0.0012
Ca 1,577 458 2,493 2,651 0.0002
Sr 1,415 458 2,331 2,513 0.0008
Ba 1,361 458 2,277 2,373 0.006

Many metals form trifluorides, such as iron, bismuth, the rare earth elements, and the metals in the aluminium column of the periodic table. At this point, the ionicity of the fluorides begins to reduce, and acidity begins to grow, even though the compounds are still ionic solids, mainly weakly basic. In many cases, while the trifluorides are still ionic, other trihalides may be volatile (for example, aluminium[237] or gold[238]). No trifluoride is soluble in water, but several are soluble in other solvents.[239]

The tetrafluorides show a mixture of ionic and covalent bonding: Zirconium, hafnium, plus many of the actinides form square-antiprismatic ionic,[240][241] high-melting[242] tetrafluorides. Titanium,[243] niobium,[244] and vanadium[245] tetrafluorides are polymeric, with melting or decomposition points below 400 °C (vanadium's may also disproportionate at 100–120 °C to the trifluoride and the pentafluoride[246]). Many transition metals tetrafluorides show unusual structures intermediate between ionic and monomeric covalent (notably iridium tetrafluoride featuring both shared and unshared fluorines[247]). They typically have low melting points (osmium tetrafluoride's is, for example, 230 °C[248]) or are of low stability (manganese tetrafluoride decomposes even at room temperature[249]). Germanium tetrafluoride forms tetrahedral molecules and is a gas at room temperature.[250]

High oxidation states

Metal penta- and higher fluorides are all covalently bonded 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 five metals (manganese heptoxide, technetium heptoxide, ruthenium tetroxide, osmium tetroxide, and iridium tetroxide[251]). Fluorine forms molecules with fifteen metals because its small size and single charge as an ion allows surrounding metal atoms with more fluorines than oxygen can.

alt1
alt2
Structure of BiF5 (α-UF5), a single chain structure and packing of chains.

Bismuth's highest fluoride is a volatile pentafluoride that is a powerful fluorinating agent. In the solid state, it is polymeric, consisting of linear chains of octahedra, sharing axial fluorides. Pentavalent bismuth behaves as an acid and forms hexafluorobismuthate, [BiF6], upon reaction with a fluoride donor, either strong (such as NaF[252][253]) or not (such as XeF4[254]), thus being an acid. Niobium and tantalum pentafluorides, while thermally stable, are also strong acids and thus hydrolyze easily. They are truly molecular, forming tetrameric molecules.[255] Many metals that form higher fluorides also can form pentafluorides, but they may be unstable. For example, osmium and ruthenium pentafluorides, while also polymeric, have a different structure, and are of low stability (hydrolyze easily, are not of the highest oxidation state, nor a stable one).[248]

The metals that make well-characterized hexafluorides include nine metals in the center of the periodic table and three actinides. At room temperature, tungsten hexafluoride is a gas, molybdenum hexafluoride and rhenium hexafluoride are liquids, and the rest are volatile solids. Metal hexafluorides are oxidants because of their tendency to get reduced: for example, platinum hexafluoride was the first compound to oxidize molecular oxygen[256] and xenon.[257] The least unstable hexafluoride is that of osmium.[248] Polonium also forms a hexafluoride, but it is understudied.[258]

test-tube like glass container, with pointy sealed end, filled with white crystals
Uranium hexafluoride in sealed glass

Rhenium is the only metal known to bond with seven fluorides in charge-neutral metal compound, which is the charged-ligands number record.[259] Rhenium heptafluoride can be synthesized from the elements, and is relatively stable against heat.[260] Chemically, it may serve as a non-oxidizing acid to form the corresponding anion, ReF
8
, or with stronger acids as a base to form a cation, ReF+
6
.[261] The molecule adopts a pentagonal bipyramid molecular geometry (expected to be shared with the currently unknown but perhaps possible iridium heptafluoride[262] (report of synthesis is being prepared[251]), technetium heptafluoride,[263] and osmium heptafluoride[264]).

Nonmetal fluorides

The nonmetal binary fluorides are volatile compounds. They show a great difference between period 2 and other fluorides. For instance, period 2 elements elements fluorides never exceed the octet in their atoms. (Boron is an exception due to its specific position in the periodic table.) Lower-period elements, however, may form hypervalent molecules, such as phosphorus pentafluoride or sulfur hexafluoride.[265] The reactivity of such species varies greatly: sulfur hexafluoride is inert, while chlorine trifluoride is extremely reactive. But there are some tendencies and trends based on periodic table columns.

Lewis dot diagram structures show three formal alternatives for describing bonding in boron monofluoride.

There are two compounds with fluorine involved in bonding with the bond order greater than 1: nitrogen monofluoride and boron monofluoride.[207] They two are isoelectronic with O2 and N2. The ground state of the BF molecule makes fluorine be the one to hold the positive charge, despite the electronegativities (although the oxidation state remains −1).[207] The NF molecule has a very small negative charge on fluorine, and both higher-energy states shift the positive charge on fluorine again.[207] The bond order of the more-studied BF molecule has been described as 1.4 (intermediate between a single and double bond). This figure comes from the resonance of the three Lewis configurations (see the image) having unequal impact on the resulting molecule.[266]

Boron trifluoride is a planar molecule. It has only six electrons around the central boron atom (and thus an incomplete octet), but it readily accepts a Lewis base, forming adducts with lone-pair-containing molecules or ions such as ammonia or another fluoride ion which can donate two more electrons to complete the octet.[267]

Silicon tetrafluoride, similar to carbon tetrafluoride and germanium tetrafluoride, adopts a molecular tetrahedral structure.[268] SiF4 is stable against heating or electric spark, but reacts with water (even moist air), metals, and alkalies, thus demonstrating weak acidic character.[269] Reactions with organomagnesium compounds, alcohols, amines, and ammonia yield adduction compounds.[269] Fluorosilicic acid, a derivative of SiF4, is a strong acid in aqueous solution (the anhydrous form does not exist).[270]

Pnictogens (nitrogen's periodic table column) show very similar trends in reactivity and acidity of the highest fluorides (pentafluorides) and most common ones (trifluorides), with the said property increasing down the group: NF3 is stable against hydrolysis,[271] PF3 hydrolyzes very slowly in moist air,[272] while AsF3 completely hydrolyzes.[271] SbF3 hydrolyzes only partially because of the increasing ionic character of the bond to fluorine. The compounds are weak Lewis bases, with NF3 again being an exception.[271] The pentafluorides of phosphorus[272] and arsenic[273] are much more reactive than their trifluorides; antimony pentafluoride is such a strong acid that it holds the title of the strongest Lewis acid.[273] The chalcogens (oxygen's periodic table column) are somewhat similar: The tetafluorides are thermally unstable and hydrolyze, and are also ready to use their lone pair to form adducts to other (acidic) fluorides. Sulfur and selenium tetrafluorides molecules are trigonal pyramids, while TeF4 is a polymer.[274] The hexafluorides are the result of direct fluorination of the elements (compare: other hexahalides of the elements do even not exist). They, like the previous group's highest fluorides, increase in reactivity with atomic number: SF6 is extremely inert, SeF6 is less noble (for example, reacts with ammonia at 200 °C (400 °F)), and TeF6 easily hydrolyzes to give an oxoacid.[274]

The period two elements, nitrogen and oxygen, fall out of line. Nitrogen is not known to form a pentafluoride, although the tetrafluoroammonium cation (NF+
4
) features nitrogen in the formal oxidation state of +5,[275] so that although its highest possible fluoride is unknown, a cationic derivative of it is known to exist. Oxygen's highest fluoride is oxygen difluoride,[274] and fluorine again can (however, for this element only theoretically as of 2012) oxidize it to a uniquely high oxidation state of +4 in a fluorocation, OF+
3
.[276]

The T-shaped (in strong accordance with the VSEPR theory) molecule of ClF3

The well-characterized heavier halogens (chlorine, bromine, and iodine) all form mono-, tri-, and pentafluorides: XF, XF3, and XF5. Of the neutral +7 species, only iodine heptafluoride is known.[277] While chlorine and bromine heptafluorides are not known, the corresponding cations ClF+
6
and BrF+
6
, extremely strong oxidizers, are.[278] Astatine is not well-studied, and although there is a report of a non-volatile astatine monofluoride,[279] its existence is debated.[280]

Many of the halogen fluorides are powerful fluorinators. Chlorine trifluoride is particularly noteworthy—readily fluorinating asbestos and refractory oxides—and may be even more reactive than chlorine pentafluoride. Used industrially, ClF3 requires special precautions similar to those for fluorine gas because of its corrosiveness and hazards to humans.[281][282]

Superacids

Several important inorganic acids contain fluorine. They are generally very strong because of the high electronegativity of fluorine. One such acid, fluoroantimonic acid (HSbF6), is the strongest charge-neutral acid known.[283] The dispersion of the charge on the anion affects the acidity of the solvated proton (in form of H
2
F+
): The compound has an extremely low pKa of −28 and is 10 quadrillion (1016) times stronger than pure sulfuric acid.[283] 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.[284]

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) pentafluoroplatinate, [XeF]+[PtF5], monofluoroxenyl(II) undecafluorodiplatinate, [XeF]+[Pt2F11], and trifluorodixenyl(II) hexafluoroplatinate, [Xe2F3]+[PtF6].[285] Bartlett's fluorination of xenon has been called one of the ten most beautiful experiments in the history of chemistry.[286] Later in 1962, xenon was reacted directly with fluorine to form the di- and tetrafluorides. Since then, chemists have made extensive efforts to form other noble gas fluorides.

black-and-white photo showing icy-looking crystals in a dish
Xenon tetrafluoride crystals

Having started the noble gases–fluorine compounds class, xenon is also the noble gas to have the most well-known such compounds. Its binary compounds include xenon difluoride, xenon tetrafluoride, and xenon hexafluoride, as well as their derivatives.[287] Xenon forms several oxyfluorides, such as xenon oxydifluoride, XeOF2, by reaction of xenon tetrafluoride with water.[288] Its upper neighbor, krypton, is the only other one to form a well-established compound: krypton difluoride. Krypton tetrafluoride was reported in 1963,[289] but was subsequently shown to be a mistaken identification; the compound seems to be very hard to synthesize now (although even the hexafluoride may exist).[290] A krypton monofluoride cation[291] has been observed.

In strong accordance with the periodic trends, radon is notably more reactive to oxidizers, including fluorine. It has been shown to readily react with fluorine to form a solid compound, which is generally thought to be radon difluoride. The exact composition is uncertain as the compound is prone to decomposition. Calculations indicate that radon difluoride may be ionic, unlike all other binary noble gas fluorides.[279]

The lighter noble gases (helium through argon) do not form stable binary fluorides. Argon forms no binary fluoride, but reacts in extreme conditions with hydrogen fluoride to form argon fluorohydride; it is the only "stable"[note 6] argon compound. Helium and neon do not form any stable chemical compounds at all. Helium forms helium fluorohydride; it has been shown to be unstable in gas phase, but there are specific conditions for which it may be stable.[292] Neon, the least reactive element,[note 7] is not now expected to form a stable compound capable of synthesis.

Ununoctium, the last currently known group 18 element, is predicted to form ununoctium difluoride, UuoF
2
, and ununoctium tetrafluoride, UuoF
4
, which is likely to have the tetrahedral molecular geometry.[295] However, only a few atoms of ununoctium have been synthesized,[296] and its chemical properties have not been examined.

Highest oxidation states: fluorine versus oxygen

Ruthenium's highest fluoride and oxide
A molecule diagram, with a blue sphere being connected with a stick to 6 yellow-green ones to form an octahedron A molecule diagram, with a blue sphere being connected with two sticks to 4 red ones to form a tetrahedron
Ruthenium hexafluoride: Six fluorines fit around the ruthenium but only make a +6 oxidation state. Ruthenium tetroxide: Four oxygens fit around the ruthenium, making a +8 oxidation state.

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.

For groups 1–5, 10, 13–16, the highest oxidation states of oxides and fluorides are always equal. Differences are only seen in chromium, groups 7–9, copper, mercury, and the noble gases. Fluorination allows elements to achieve relatively low[note 8] oxidation states that are, however, hard to achieve. For example, no binary oxide is known for krypton, but krypton difluoride is well-studied.[297] At the same time, very high oxidation states are known for oxygen-based species only. For the previously mentioned volatile oxides, there are no corresponding hepta- or octafluorides. (For example, ruthenium octafluoride is unlikely to be ever synthesized,[263] while ruthenium tetroxide has even found an industrial use.[298]) 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.[299][note 9] However, octafluoride anions, such as the octafluoroiodate (IF
8
), octafluorozirconate (ZrF4−
8
), and octafluoroxenate (XeF2−
8
) anions are well-known.

The highest oxidation states may be uncommon to everyday life, or even industrial usage. For example, the synthesis of mercury tetrafluoride, the first compound to achieve an oxidation state above +2 for a group 12 element, breaking the filled 5d-shell, again showing the significance of the relativistic effects on the heavy elements, and fueling the debate over whether mercury, cadmium, and zinc are transition metals,[301] occurred at cryogenic temperatures and the compound decomposes at the temperatures of solid nitrogen.[302] More unstable still, the only cobalt(V) species, the CoF+
4
cation, has only been observed in gas phase (with no interactions with other atoms, thus no shown stability in any chemical environment).[263] The reason why such unstable species exist is complicated, yet can be summarized as follows on the example of the hypothesized NF
5
molecule: According to the modern calculations, five fluorine atoms and one nitrogen atoms can theoretically arrange themselves in different ways, such as NF
3
and F
2
, NF
4
and F
, NF
5
, etc. The NF
3
+F
2
system is of the smallest energy (most stable). However, if a NF
5
molecule was synthesized, it would have to go through a high-energy transitional state, from which it could decay into two molecules. But since the transitional state is higher in energy than the hexatomic molecule, the energy difference would need to be added to reach the transitional state and thus allow the decay. This energy is called reaction activation barrier. (The second decay mode has an analogical position.) Thus if little energy was added (low temperatures), then the compound could exist; however, the synthesis is a serious problem (not yet solved).[303]

Organic

Organofluorine compounds are defined by having a carbon–fluorine chemical bond. This bond is the strongest bond in organic chemistry and is very stable.[304] After the reaction, the molecular size is not changed significantly, as the elements have very similar van der Waals radii.[305] Organofluorines are very stable; there are multiple reasons contributing to this. One is that all perfluoro species twist themselves, so that the carbon chain is wrapped with the electronegatives fluorines; low polarizability and high bond energy are also important.[305]

The carbon–fluorine bond is formed in three principal ways: Fluorine replaces a halogen or hydrogen, or adds across a multiple bond. The direct reaction of hydrocarbons with fluorine gas can be dangerously reactive (the temperature may need to be lowered even to −150 °C (−240 °F)[306], so fluorinating reagents (collectively called "solid fluorine carriers", they can release fluorine upon heating, notably cobalt trifluoride[307]) may be used instead, or hydrogen fluoride. Direct fluorination becomes even less important when it comes to organohalogens or unsaturated compounds reactions, or when a prefluorocarbon is desired (then HF-based electrolysis is typically used).[308] The polymers are formed by polymerizing free radicals; other techniques used for hydrocarbon polymers do not work in that way with fluorine.[309] Mild fluorination of polyethylene, however, gives a polymer similar to the original, but with a fluorination surface (0.01 mm at maximum). This is somewhat similar to metal passivation: the bulk properties are not affected, but the surface properties are, most notably, a greater vapor barrier.[310]

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.[81]

Some important fluorocarbons
Fluoromethane (methyl
fluoride), used in
semiconductor processing
Tetrafluoroethane
(R-134a), a HFC
Dichlorodifluoromethane
(R-12 refrigerant), a CFC
Trifluoromethanesulfonic
acid
(triflic acid),
a superacid
Perfluorooctanesulfonic
acid
, a fluorosurfactant
(the anion is depicted)
A section of
polytetrafluoroethylene
(Teflon) polymer

Small molecules

Monofluoroalkanes (alkanes with one hydrogen replaced with fluorine) may be chemically and thermally unstable, yet are soluble in many solvents; but as more fluorines are in instead of hydrogens, the stability increases, while melting and boiling points, and solubility decrease. While the densities and viscosities are increased, the dielectric constants, surface tensions, and refractive indices fall.[311]

Partially fluorinated alkanes are the hydrofluorocarbons (HFCs). Substituting other halogens in combination with fluorine gives rise to chlorofluorocarbons (CFCs) or bromofluorocarbons (BFCs) and the like (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.[312] As with fluorocarbons, chlorofluorocarbons and bromofluorocarbons are not flammable: they do not have carbon–hydrogen bonds to react and released halides quench flames.[312]

two six-membered rings joined on a side, all F substituted
Perfluorodecalin (the trans isomer is shown), a perfluorocarbon that is a liquid at room temperature. It boils at a lower temperature than its hydrocarbon analog, decalin.

When all hydrogens are replaced with fluorine to achieve perfluoroalkanes, a great difference is revealed. Such compounds are extremely stable, and only sodium in liquid ammonia attacks them at standard conditions. They are also very insoluble, with few organic solvents capable of dissolving them.[311]

However, if a perfluorocarbon contains double or triple bonds (perfluoroalkenes or -alkynes), a very reactive towards ligand accepting result, even less stable than corresponding hydrocarbons.[313] Difluoroacetylene, which decomposes even under liquid nitrogen temperatures,[314] is a notable example. If such a molecule is asymmetric, then the more fluorinated carbon is attacked, as it holds positive charge caused by the C–F bonds and is shielded weakly[313] (similarly to that how unsaturated hydrocarbons attacked by HF add hydrogen to the more hydrogen-rich atom per Markovnikov's rule[315]).

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 (even though another definition exists[316]). They share most of perfluorocarbon properties (inertness, stability, non-wettingness and insolubility in water and oils, slipperiness, etc.),[317] but may differ because of the functional group properties, although the perfluorocarbon tail differ the group-specific properties as compared to those of hydrocarbon-tailed compounds.

Perfluoroalkanic acids organic acids may also be seen as perfluorinated compounds, with perfluoroalkyl joined to a carboxyl group. As the fluorines are added to the acid, its strength grows: consider acetic acid and its mono-, di-, and trifluoroacetic derivatives and their pKa values (4.74, 2.66, 1.24, and 0.23).[318] This happens because with fluorines, the anion formed after giving the proton off becomes stable, an effect caused by fluorine's great inductive effect. Because of this, the trifluoro derative is 33,800 times stronger an acid than acetic.[319] Similarly, the acidity is greatly increased for other perfluorocarboxyl acids, as well as the amines (here, it makes more sense to talk about elimination of basicity, though).[305]

The perfluoroalkanesulfonic acids are also very notable for their acidity. The sulfonic acid derivative, trifluoromethanesulfonic acid, is comparable in strength to perchloric acid.[320] These compounds lower surface energy; for this reason, they, especially perfluorooctanesulfonic acid (PFOS, formerly the active component in brand "Scotchgard") have found industrial use as surfactants (see above).[320]

If a perfluorinated compound has a fluorinated tail, but also a few non-fluorinated carbons (typically two) near the functional group, it is called a fluorotelomer (such molecules are commercially treated as perfluorinated), but such molecules are more of industrial value than chemical. The chain end may similarly be attached to different functiuonal groups (via the hydrogenized terminal carbon), such as hydroxyl resulting in fluorotelomer alcohols, sulfonate resulting in fluorotelomer sulfonates, etc.[317]

Polymers

Fluoropolymers are similar in many regards with smaller molecules; adding fluorine to a polymer affects the properties in the same manner as in small molecules (increasing chemical stability, melting point, reducing flammability, solubility, etc.). Each fluoropolymer has own characteristic properties, though.[309]

Lines (wire-like), messed up (A) and almost linear (B)
(A) is an unstressed elastomer; (B) is the same elastomer under stress; it will return to the A configuration with stress removed. (The dots represent cross-links)

The simplest fluoroplastic is polytetrafluoroethylene (PTFE, DuPont brand Teflon), which 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. Its resistance to van der Waals forces makes PTFE the only known surface to which a gecko cannot stick.[321] The compound, however, lacks an ability to transform upon melting, which is not a problem for various PTFE deratives, namely FEP (fluorinated ethylene propylene, with some fluorines replaced with the –CF3 group) or PFA (perfluoroalkoxy, some fluorines replaced with –OCF3). They share most properties with PTFE, but there are still differences, namely maximum usage temperature (highest for the non-flexible PTFE).[322]

There are other fluoroplastics other than perfluorinated. A notable example is PVF (polyvinyl fluoride, structural unit: –CH2CHF–), with one one-fourth of fluorines compared to those PTFE. Despite this, it still has many properties of more fluorinated compounds.[323] PCTFE (polychlorotrifluoroethylene, structural unit: –CF2CFCl–) is another important compound. It differs from PTFE by having a quarter of fluorine replaced with chlorines, yet this difference brings even greater hardness, creep resistance, and moisture persistence.[323]

Fluoroelastomers are another class of fluoropolymers. Elastomers are polymers which can become much longer under stress and contract back when the stress is removed. Non-crosslinked fluoroelastomers are viscous incompressible liquids, and they do not have the said properties.[324] Crosslinks are essential: they "fix" a given configuration, and while the connected chains can lengthen toward an external force, they pull themselves back with the force removed.[325] However, usual crosslinked perfluoropolymers do not exhibit such elastomeric properties at room temperature and below.[324] The happens because the true elastomers have a high number of "realizations of the system" (degree of disorder or entropy), which lowers upon straightening, increasing the energy of the system, making this state unfavorable.[325] In contrast, crosslinked perfluorocarbons cannot revolve around the carbon skeleton (lack alternative conformations).[324] High fluorine content present is again helpful for solvent resistance and thermal stability.[324]

Skeletal chemical formula
The complex unit structure of Nafion

Nafion is a structurally complicated polymer. It has a PTFE-like backbone, but also contains side chains of perfluoro ether that end in sulfonic acid (–SO2OH) groups. It also possesses great chemical stability, while exact properties vary with morphology. However, because of the difficult chemical structure, it is also relatively easily converted to an ionomer (shows conductivity) by adding cations like Na+ or by converting into the sulfonic acid rather than the given sulfonyl fluoride. The conductivity is due to that the main carbon chain separates from the side chains, thus forming polar and non-polar regions. This form is also very hydroscopic.[326]

See also

  • Halogens—the periodic group to which fluorine belongs
  • Caesium—the least electronegative element, fluorine's "opposite"

Notes

118 squares with numbers and letters in them, mostly colored gray and green, with a few numbers and words outside the boxes
Periodic table colored to show how elements are treated in this article: dark gray elements are metals, green ones are nonmetals, blue ones are the noble gases, the purple one is hydrogen, the yellow one is carbon, and the light gray ones are those with indeterminate properties.
  1. ^ Density of air at 100 kilopascal and 0 °C is 1.2724 grams per liter.[17]
  2. ^ It differs from this value in elemental fluorine, where the atoms are bonded to each other and thus at oxidation state 0, and a few polyatomic ions: The very unstable anions F
    2
    and F
    3
    with intermediate oxidation states exist at very low temperatues, decomposing at around 40 K.[205] Also, the F+
    4
    cation and a few related species have been predicted to be stable.[206]
  3. ^ 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 is included for comparison among nonmetals, even though it is closer to metals chemically than to nonmetals. The noble gases are treated separately from nonmetals; hydrogen is discussed in the Hydrogen fluoride section and carbon in the Organic compounds 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, the yellow one is carbon, and the light gray elements have unknown properties.
  4. ^ See citation for more detailed explanation.[214]
  5. ^ Energies are given in −kJ/mol, solubilities in mol/L; HE means "hydration energy".
  6. ^ In the context of the noble-gas compounds, "stable" means not decaying chemically over time under certain (any count) conditions; other ("meta-stable") compounds are not treated in this overview article.
  7. ^ Even though helium and neon have no known stable compounds, calculations show that helium may be called slightly less unreactive than neon (may, unlike neon, form some neutral-charged compounds that will not decay over time in specific conditions, and a stable cation of helium (HeH+, the strongest acid of all species,[293] yet existent) has been known since 1925, while no neon cations are. Additionally, clathrates are known for every noble gas but neon.[294]
  8. ^ 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 "high" here, but for larger palladium and platinum +6 is still "low".
  9. ^ 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 completely ionic because of large sixth, seventh, and eighth oxidation energies of all elements, which make the reactions that may produce such compounds highly unfavorable (even though the possible osmium octafluoride may be mostly ionic[300]).

References

  1. ^ "Standard Atomic Weights: Fluorine". CIAAW. 2021.
  2. ^ Prohaska, Thomas; Irrgeher, Johanna; Benefield, Jacqueline; Böhlke, John K.; Chesson, Lesley A.; Coplen, Tyler B.; Ding, Tiping; Dunn, Philip J. H.; Gröning, Manfred; Holden, Norman E.; Meijer, Harro A. J. (4 May 2022). "Standard atomic weights of the elements 2021 (IUPAC Technical Report)". Pure and Applied Chemistry. doi:10.1515/pac-2019-0603. ISSN 1365-3075.
  3. ^ a b c d Jaccaud et al. 2000, p. 381.
  4. ^ a b c Haynes 2011, p. 4.121.
  5. ^ a b Jaccaud et al. 2000, p. 382.
  6. ^ a b c Compressed Gas Association 1999, p. 365.
  7. ^ "Triple Point | The Elements Handbook at KnowledgeDoor". KnowledgeDoor.
  8. ^ Himmel, D.; Riedel, S. (2007). "After 20 Years, Theoretical Evidence That 'AuF7' Is Actually AuF5·F2". Inorganic Chemistry. 46 (13). 5338–5342. doi:10.1021/ic700431s.
  9. ^ Dean 1999, p. 4.6.
  10. ^ Dean 1999, p. 4.35.
  11. ^ Matsui 2006, p. 257.
  12. ^ Yaws & Braker 2001, p. 385.
  13. ^ a b Mackay, Mackay & Henderson 2002, p. 72.
  14. ^ Cheng et al. 1999.
  15. ^ Chisté & Bé 2011.
  16. ^ a b Jassaud et al. 2005, p. 2.
  17. ^ Shelquist, Richard (2010). "An introduction to air density and density altitude calculations". Shelquist Engineering. Retrieved 29 April 2011.
  18. ^ a b c Greenwood & Earnshaw 1998, p. 801.
  19. ^ Burdon, J.; Emson, B.; Edwards, A. J. (1987). "Is fluorine really yellow?". Journal of Fluorine Chemistry. 34 (3–4): 471–474. doi:10.1016/S0022-1139(00)85188-X.
  20. ^ a b c Hemat, R. A. S. (2009). Water. Vol. 42. Urotext. pp. 707–709. ISBN 9781903737125.
  21. ^ a b Dean 1999, p. 523.
  22. ^ Cite error: The named reference phases was invoked but never defined (see the help page).
  23. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1063/1.1711946, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1063/1.1711946 instead.
  24. ^ a b Jassaud et al. 2005, p. 1.
  25. ^ 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): 2832–2838. doi:10.1039/b801115j.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  26. ^ Dean 1999, p. 564.
  27. ^ Lide 2004, p. 10-137–10-138.
  28. ^ Cheng, H.; Fowler, D. E.; Henderson, P. B.; Hobbs, J. P.; Pascaloni, M. R. (1999). "On the magnetic susceptibility of fluorine". Journal of Physical Chemistry A. 103: 2861–2866.
  29. ^ a b Template:Ru icon Akhmetov, N. S. (2001). Общая и неорганическая химия (4th ed.). p. 312. ISBN 5-06-003363-5. {{cite book}}: Unknown parameter |trans_title= ignored (|trans-title= suggested) (help)
  30. ^ Moore, John W.; Stanitski, Conrad L.; Jurs, Peter C. (2010). Principles of chemistry: The molecular Science. Cengage Learning. p. 156. ISBN 978-0-495-39079-4. Retrieved 7 May 2011.
  31. ^ Greenwood & Earnshaw 1998, p. 804.
  32. ^ Macomber, Roger S. (1996). Organic chemistry. Vol. 1. University Science Books. p. 230. ISBN 0-935702-90-3. Retrieved 26 July 2011.
  33. ^ Nelson, Eugene W. (1947). "'Bad man' of the elements". Popular Mechanics. 88 (2): 106–108, 260.
  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): 3979–3983. 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. {{cite journal}}: |access-date= requires |url= (help)
  36. ^ Kahn, Bernd (2007). Radioanalytical chemistry. ISBN 978-0-387-34122-4. Retrieved 4 May 2011.
  37. ^ Even, J.; et al. (2008). First transactinide chemistry behind TASCA (PDF) (Report). Gesellschaft für Schwerionenforschung. p. 1. Retrieved 29 April 2011. {{cite report}}: Explicit use of et al. in: |author= (help)
  38. ^ Lidin, Molochko & Andreeva 2000, pp. 442–455.
  39. ^ Greenwood & Earnshaw 1998, p. 844.
  40. ^ Jassaud et al. 2005, p. 3.
  41. ^ Pitzer, Kenneth S. (1975). "Fluorides of radon and element 118". Journal of the Chemical Society: Chemical communications: 760b–761. doi:10.1039/C3975000760B.
  42. ^ Khriachtchev, Leonid; Pettersson, Mika; Runeberg, Nino; Lundell, Jan; Räsänen, Markku (2000). "A stable argon compound". Nature. 406 (6798): 874–876. doi:10.1038/35022551. PMID 10972285. Retrieved 29 April 2011.
  43. ^ Lidin, Molochko & Andreeva 2000, p. 252.
  44. ^ National Nuclear Data Center. "NuDat 2.1 database – Fluorine-19". Brookhaven National Laboratory. Retrieved 31 September 2005. {{cite web}}: Check date values in: |accessdate= (help)
  45. ^ a b National Nuclear Data Center. "NuDat 2.1 database". Brookhaven National Laboratory. Retrieved 31 September 2005. {{cite web}}: Check date values in: |accessdate= (help)
  46. ^ National Nuclear Data Center. "NuDat 2.1 database – Fluorine-18". Brookhaven National Laboratory. Retrieved 31 September 2005. {{cite web}}: Check date values in: |accessdate= (help)
  47. ^ a b Audi, G.; Wapstra, A. H.; Thibault, C.; Blachot, J.; Bersillon, O. (2003). "The NUBASE evaluation of nuclear and decay properties" (PDF). Nuclear Physics A. 729: 3–128. Bibcode:2003NuPhA.729....3A. doi:10.1016/j.nuclphysa.2003.11.001.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  48. ^ Cameron, A. G. W. (1973). "Abundance of the elements in the Solar System" (PDF). Space Science Review. 15: 121–146. Bibcode:1973SSRv...15..121C. doi:10.1007/BF00172440.
  49. ^ a b Croswell, Ken (2003). "Fluorine: An element–ary mystery". Sky and Telescope. Retrieved 3 May 2011.
  50. ^ a b =Renda, A.; et al. (2004). "On the origin of fluorine in the Milky Way" (PDF). Monthly Notices of the Royal Astronomical Society. 354 (2): 575–580. arXiv:astro-ph/0410580. Bibcode:2004MNRAS.354..575R. doi:10.1111/j.1365-2966.2004.08215.x. {{cite journal}}: Explicit use of et al. in: |author= (help)CS1 maint: extra punctuation (link) CS1 maint: unflagged free DOI (link)
  51. ^ Neufeld, David; Bergin, Edwin; Gerin, Maryvonne (2010). "Tracing the Milky Way's hidden reservoirs of gas". European Space Agency. {{cite web}}: |access-date= requires |url= (help); Missing or empty |url= (help)CS1 maint: multiple names: authors list (link)
  52. ^ 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.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  53. ^ Snow, Theodore, P. (2007). "The abundance of interstellar fluorine and its implications" (PDF). The Astrophysical Journal. 655 (1): 285–298. arXiv:astro-ph/0611066. Bibcode:2007ApJ...655..285S. doi:10.1086/510187. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)CS1 maint: multiple names: authors list (link)
  54. ^ Zhang, Y.; Liu, X.-W. (2005). "Fluorine abundances in planetary nebulae" (PDF). The Astrophysical Journal. 631 (1): L61–L63. arXiv:astro-ph/0508339. Bibcode:2005ApJ...631L..61Z. doi:10.1086/497113. Retrieved 3 May 2011.
  55. ^ Jassaud et al. 2005, p. 4.
  56. ^ a b c d e Greenwood & Earnshaw 1998, p. 795.
  57. ^ Kelly, T.D. "Historical fluorspar statistics" (PDF). U.S. Geological Survey Data Series 140. United States Geological Service. Retrieved 25 January 2012.
  58. ^ Lusty, P. A. J. (2008). "The need for indigenous fluorspar production in England". British Geological Survey. Retrieved 25 January 2012. {{cite web}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  59. ^ a b Norwood, Charles J. (1907). "Fluorspar and its Occurrence". Kentucky geological survey Bulletin 9: Fluorspar deposits of Kentucky. Globe Printing Company. p. 52. {{cite book}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  60. ^ a b Gribble, Gordon W. (2002). "Naturally occurring organofluorines". The Handbook of Environmental Chemistry. 3N: 121–136. doi:10.1007/10721878_5.
  61. ^ Template:De icon Schmedt, Jörn; Mangst, Martin; Kraus, Florian (2012). "Elementares Fluor F2 in der Natur – In-situ-Nachweis und Quantifizierung durch NMR-Spektroskopie". Angewandte Chemie. 124 (31): 7968–7971. doi:10.1002/ange.201203515.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  62. ^ Greenwood & Earnshaw 1998, p. 790.
  63. ^ Senning, Alexander (2007). Elsevier's dictionary of chemoetymology: The whies and whences of chemical nomenclature and terminology. Elsevier. p. 149. ISBN 978-0-444-52239-9. Retrieved 7 May 2011.
  64. ^ Agricola, Georgius (1912). De Re Metallica. trans. Hoover, Herbert Clark; Hoover, Henry Lou. pp. preface, 380–381 (3, 416–417 in linked viewer).
  65. ^ Greenwood & Earnshaw 1998, p. 109.
  66. ^ a b c Asimov, Isaac (1966). The noble gases. Basic Books. p. 162. ISBN 978-0-465-05129-8.
  67. ^ a b Peter, Meiers (2010). "Discovery of fluorine". Fluoride History. Retrieved 17 January 2011.
  68. ^ Partington, J. R. (1923). "The early history of hydrofluoric acid". Memoirs and Proceedings of the Manchester Literary and Philosophical Society. 67 (6): 73–87.
  69. ^ a b c d e f g h Kirsch, Peer (2004). "Fluorine". Modern fluoroorganic chemistry: Synthesis, reactivity, applications. pp. 3–10. ISBN 978-3-527-30691-6. Retrieved 7 May 2011.
  70. ^ Template:Fr icon Ampere, André-Marie (1816). "Suite d'une classification naturelle pour les corps simples". Annales de chimie et de physique: 1–5. Retrieved 7 May 2011.
  71. ^ a b Weeks, Mary Elvira (1932). "The discovery of the elements. XVII. The halogen family". Journal of Chemical Education. 9 (11): 1915–1939. Bibcode:1932JChEd...9.1915W. doi:10.1021/ed009p1915.
  72. ^ "09 Fluorine". Elements.vanderkrogt.net. Retrieved 24 January 2012.
  73. ^ a b c d e f Storer, Frank Humphreys (1864). First outlines of a dictionary of solubilities of chemical substances. Cambridge. pp. 278–280. ISBN 978-1-176-62256-2.
  74. ^ "Element 114 is named flerovium and element 116 is named livermorium". iupac.org. IUPAC. 2012. Retrieved 1 June 2012.
  75. ^ Template:Fr icon Moissan, Henry (1886). "Action d'un courant électrique sur l'acide fluorhydrique anhydre". Comptes rendus hebdomadaires des séances de l'Académie des sciences. 102: 1543–1544. Retrieved 7 May 2011.
  76. ^ Greenwood & Earnshaw 1998, pp. 789–791.
  77. ^ "The Nobel Prize in chemistry 1906". Nobelprize.org. Retrieved 7 July 2009.
  78. ^ Greenwood & Earnshaw 1998, p. 791.
  79. ^ James, Laylin K. (1993). Nobel laureates in chemistry, 1901–1992. Chemical Heritage Foundation. p. 35. ISBN 978-0-8412-2690-6.
  80. ^ Hounshell & Smith 1988, p. 156.
  81. ^ a b c d 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): 276–289. Bibcode:2009PJAB...85..276O. doi:10.2183/pjab.85.276.
  82. ^ "Freon". DuPont.com. Retrieved 10 November 2012.
  83. ^ Hounshell & Smith 1988, p. 157.
  84. ^ Meyer, Eugene (1977). Chemistry of hazardous materials (3 ed.). p. 111. ISBN 9780131292390. {{cite book}}: |journal= ignored (help)
  85. ^ Hounshell & Smith 1988, p. 489.
  86. ^ Harrington, Ann (2003). "Who's afraid of a new product? Not W.L. Gore. It has mastered the art of storming completely different businesses". CNN Money (Fortune Magazine). Retrieved 21 January 2012.
  87. ^ US EPA (2008). "Ozone depletion glossary". Retrieved 3 September 2008.
  88. ^ "Brief questions and answers on ozone depletion | Ozone layer protection | US EPA". Epa.gov. 2006. Retrieved 8 November 2011.
  89. ^ "Freedonia industry study #1555 – Fluorochemicals" (PDF). Freedonia. Retrieved 23 June 2012.
  90. ^ a b c "Global fluorochemicals Market to exceed 2.6 million tons by 2015, according to a new report by Global Industry Analysts, Inc". Global Industry Analysts (via PRWeb). 2010. Retrieved 26 January 2012.
  91. ^ "Fluorochemical develops rapidly in China". China Chemical Reporter (Goliath). 2002. Retrieved 26 January 2012.
  92. ^ a b Kogelfirst2= Nikhil C.; Trivedi; Barker, James M. (2006). Industrial minerals & rocks: Commodities, markets, and uses. Society for Mining, Metallurgy, and Exploration (U.S.). pp. 461–473. ISBN 978-0-87335-233-8.{{cite book}}: CS1 maint: numeric names: authors list (link)
  93. ^ a b c d e f g h i j k l m 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.
  94. ^ Miller, M. Michael (2003). "Fluorspar". U.S. Geological Survey. pp. 27.1–27.12. {{cite book}}: Missing or empty |title= (help); Unknown parameter |name= ignored (help)
  95. ^ Aigueperse et al. 2005, p. 33.
  96. ^ Aigueperse et al. 2005, pp. 25–26.
  97. ^ Willey, Ronald R. (2007). Practical equipment, materials, and processes for optical thin films. Willey Optical. p. 113. ISBN 9780615143972.
  98. ^ DOE/Lawrence Berkeley National Laboratory (2009). "'Invisibility cloak' successfully hides objects placed under it". Science Daily. Retrieved 31 January 2012.
  99. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1038/nmat2461, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1038/nmat2461 instead.
  100. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1038/nnano.2011.82, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1038/nnano.2011.82 instead.
  101. ^ Aigueperse et al. 2005, pp. 26–27.
  102. ^ Committee on Assessment of Fire Suppression Substitutes and Alternatives to Halon (Commission on Physical Sciences, Mathematics, and Applications, National Research Council) (1997). "Executive summary". Fire suppression substitutes and alternatives to Halon for U.S. Navy applications. National Academies Press. p. 1. ISBN 978-0-309-07492-6.{{cite book}}: CS1 maint: multiple names: authors list (link)
  103. ^ "Fluoropolymers to 2013 – Demand and sales forecasts, market share, market size, market leaders (Abstract for market report)". Freedonia Group. 2012. Retrieved 10 November 2012.
  104. ^ a b Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1134/S1070363209030335, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1134/S1070363209030335 instead.
  105. ^ a b c d e Martin, John Wilson (2007). Concise encyclopedia of the structure of materials. Elsevier. pp. 187–194. ISBN 978-0-08-045127-5.
  106. ^ Nakagawa, Ulara (2011). "15 sights that make Tokyo so fascinating". CNN. Retrieved 31 December 2011.
  107. ^ Bhiwankar, Nikhil (2011). "Weathering the storm: Fluoropolymer films protect solar modules and ensure performance". Altenergymag.com. Retrieved 31 December 2011.
  108. ^ DeBergalis, Michael (2004). "Fluoropolymer films in the photovoltaic industry" (PDF). Journal of Fluorine Chemistry. 125: 1255–1257. doi:10.1016/j.jfluchem.2004.05.013.
  109. ^ Grot, Walter (2011). Fluorinated ionomers. Elsevier. pp. 1–10. ISBN 978-1-4377-4457-6.
  110. ^ Ramkumar, Jayshreee (2012). "Nafion Persulphonate Membrane: Unique Properties and Various Applications". In Banerjee, S. (ed.). Functional materials: Preparation, processing and applications. Elsevier. pp. 549–578. ISBN 978-0-12-385142-0.
  111. ^ Burney, H.S. (1999). "Past, Present, and Future of the Chlor-Alkali Industry". Chlor-alkali and chlorate technology: R.B. MacMullin memorial symposium. The Electrochemical Society. pp. 105–126. ISBN 1-56677-244-3.
  112. ^ a b Renner, R. (2006). "The long and the short of perfluorinated replacements". Environmental Science & Technology. 40 (1): 12. doi:10.1021/es062612a. PMID 16433328.
  113. ^ Kissa, Erik (2001). Fluorinated surfactants and repellents. Marcel Dekker. pp. 516–551. ISBN 0-8247-0472-X.
  114. ^ Ullman, Fritz (2008). Ullmann's fibers: Textile and dyeing technologies, high performance and optical fibers. Vol. 2. Wiley-VCH. pp. 538, 543–547. ISBN 978-3-527-31772-1.
  115. ^ a b c d Jassaud et al. 2005, p. 12.
  116. ^ Aigueperse et al. 2005, p. 35.
  117. ^ Jassaud et al. 2005, pp. 11–12.
  118. ^ El-Kareh, Badih (1994). "Fluorine –Based Plasmas". Fundamentals of semiconductor processing technology. p. 317. ISBN 978-0-7923-9534-8. Retrieved 7 May 2011.
  119. ^ 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.
  120. ^ Krieger, F. J. (1960). The Russian literature on rocket propellant (PDF) (Report). The Rand Corporation. p. 17. Retrieved 31 February 2011. {{cite report}}: Check date values in: |accessdate= (help)
  121. ^ Sutton, Oscar; Biblarz (2010). "Liquid Oxidizers". Rocket propulsion elements. p. 256. ISBN 978-0-470-08024-5. Retrieved 7 May 2011.
  122. ^ a b Jassaud et al. 2005, pp. 4–5.
  123. ^ Jassaud et al. 2005, p. 6.
  124. ^ a b Jassaud et al. 2005, pp. 10–11.
  125. ^ Shriver, Duward; Atkins, Peter (2010). Solutions manual for inorganic chemistry. Macmillan. p. 427. ISBN 978-1-4292-5255-3.
  126. ^ Kirsch, Peer (2004). Modern fluoroorganic chemistry: Synthesis, reactivity, applications. John Wiley & Sons. p. 7. ISBN 3-527-30691-9.
  127. ^ Christe, K. (1986). "Chemical synthesis of elemental fluorine". Inorganic Chemistry. 25 (21): 3721–3724. doi:10.1021/ic00241a001. {{cite journal}}: Unknown parameter |years= ignored (help)
  128. ^ Olivares, M.; Uauy, R. (2004). Essential nutrients in drinking water (Draft) (PDF) (Report). WHO. Retrieved 30 December 2008.
  129. ^ 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.
  130. ^ Pizzo G.; Piscopo, M. R.; Pizzo, I.; Giuliana, G. (2007). "Community water fluoridation and caries prevention: a critical review" (PDF). Clinical Oral Investigation. 11 (3): 189–193. doi:10.1007/s00784-007-0111-6. PMID 17333303.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  131. ^ Centers for Disease Control and Prevention (2001). "Recommendations for using fluoride to prevent and control dental caries in the United States". MMWR Recommendations and Reports. 50 (RR-14): 1–42. PMID 11521913.
  132. ^ Ripa, L. W. (1993). "A half-century of community water fluoridation in the United States: review and commentary" (PDF). Journal of Public Health Dentistry. 53 (1): 17–44. doi:10.1111/j.1752-7325.1993.tb02666.x. PMID 8474047.
  133. ^ a b Cheng, K. K.; Chalmers, I.; Sheldon, T. A. (2007). "Adding fluoride to water supplies" (PDF). BMJ. 335 (7622): 699–702. doi:10.1136/bmj.39318.562951.BE. PMC 2001050. PMID 17916854.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  134. ^ Marya, C. M. (2011). A textbook of public health dentistry. JP Medical Limited. p. 343. ISBN 9789350252161. {{cite book}}: |access-date= requires |url= (help)
  135. ^ Armfield, J. M. (2007). "When public action undermines public health: A critical examination of antifluoridationist literature". Australia and New Zealand Health Policy. 4 (1): 25. doi:10.1186/1743-8462-4-25. PMC 2222595. PMID 18067684.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  136. ^ a b Fejerskov, Ole (2008). Dental caries: The disease and its clinical management. John Wiley & Sons. p. 518. ISBN 978-1-4051-3889-5. {{cite book}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)CS1 maint: extra punctuation (link)
  137. ^ National Health and Medical Research Council (Australia) (2007). "A systematic review of the efficacy and safety of fluoridation" (PDF). Retrieved 24 February 2009. Summary: Yeung, C. A. (2008). "A systematic review of the efficacy and safety of fluoridation". Evidence-Based Dentistry. 9 (2): 39–43. doi:10.1038/sj.ebd.6400578. PMID 18584000. {{cite journal}}: Unknown parameter |laydate= ignored (help); Unknown parameter |laysource= ignored (help); Unknown parameter |laysummary= ignored (help)
  138. ^ Cracher, Connie Myers (2009). "Current concepts in preventive dentistry" (PDF). dentalcare.com. p. 12. Retrieved 20 January 2012.
  139. ^ 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 978-0-309-10128-8.
  140. ^ Swinson, Joel (2005). "Fluorine – A vital element in the medicine chest" (PDF). Pharmaceutical Chemistry. pp. 26–27. Retrieved 26 August 2010. {{cite web}}: Unknown parameter |jornal= ignored (help)
  141. ^ Schubiger, P. A. (2006). Pet chemistry: The driving force in molecular imaging. Springer. p. 144. ISBN 9783540326236.
  142. ^ Emsley, John (2011). Nature's building blocks: An A-Z guide to the elements (2nd ed.). Oxford University Press. p. 178. ISBN 978-0-19-960563-7.
  143. ^ Raj, P. Prithvi; Erdine, Serdar (2012). Pain-relieving procedures: The illustrated guide. John Wiley & Sons. p. 58. ISBN 9781118300459.{{cite book}}: CS1 maint: multiple names: authors list (link)
  144. ^ Filler, R.; Saha, R. (2009). "Fluorine in medicinal chemistry: A century of progress and a 60-year retrospective of selected highlights" (PDF). Future Medicinal Chemistry. 1 (5): 777–791. doi:10.4155/fmc.09.65. PMID 21426080.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  145. ^ Mitchell, Siobhan E.; Triggle, D. J. (2004). Antidepressants. Infobase Publishing. pp. 37–39. ISBN 978-1-4381-0192-7.{{cite book}}: CS1 maint: multiple names: authors list (link)
  146. ^ 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" (PDF). Clinical Infectious Diseases. 44 (7): 977–980. doi:10.1086/512369. PMID 17342653.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  147. ^ 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 (PDF) (Report). U.S. Geological Survey Scientific Investigations. pp. 7–8. Retrieved 6 June 2011.{{cite report}}: CS1 maint: multiple names: authors list (link)
  148. ^ "Fluorine's treasure trove". ICIS news. 2006. Retrieved 20 February 2011.
  149. ^ Theodoridis, George (2006). Fluorine and the Environment: Agrochemicals, Archaeology, Green Chemistry & Water. Elseiver. pp. 121–176. ISBN 9780444526724.
  150. ^ Eisler, Ronald (1995). Biological report 27: Sodium monofluoroacetate (1080) Hazards to fish, wildlife and invertebrates: A synoptic review (PDF) (Report). Patuxent Environmental Science Center (U.S. National Biological Service). Retrieved 5 June 2011.
  151. ^ a b c Proudfoot, A. T.; Bradberry, S. M.; Vale, J. A. (2006). "Sodium fluoroacetate poisoning". Toxicology Review. 25 (4): 213–219. doi:10.2165/00139709-200625040-00002. PMID 17288493.{{cite journal}}: CS1 maint: multiple names: authors list (link) Cite error: The named reference "Proudfoot" was defined multiple times with different content (see the help page).
  152. ^ "Class I ozone-depleting substances". Sodium fluoride – pesticidal uses. Scorecard. Retrieved 20 February 2011.
  153. ^ Siegemund et al. 2005, p. 30.
  154. ^ "Fact sheet: Trifluralin". Pesticides News. 52: 20–21. 2001.
  155. ^ European Commission (2007). Trifluralin (PDF) (Report). {{cite report}}: Cite has empty unknown parameter: |1= (help)
  156. ^ Case T-475/07, Dow AgroSciences Ltd vs. European Commission (2011). The General Court of European Union (Third Camber).
  157. ^ Barnette, William E. (1995). "Physcial Organic Aspects of Fluorinated Argichemicals". Fluorine in agriculture. Smithers Rapra Publishing. pp. 1–19. ISBN 9781859570333.
  158. ^ a b Murphy, C.; Schaffrath, C.; O'Hagan, D. (2003). "Fluorinated natural products: The biosynthesis of fluoroacetate and 4-fluorothreonine in Streptomyces cattleya" (PDF). Chemosphere. 52 (2): 455–461. doi:10.1016/S0045-6535(03)00191-7. PMID 12738270.
  159. ^ 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.
  160. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1039/c0em00373e, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1039/c0em00373e instead.
  161. ^ 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" (PDF). Journal of spinal disorders. 13 (6): 541–544. doi:10.1097/00002517-200012000-00016. PMID 11132989.
  162. ^ 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): 3735–3739. Bibcode:1977PNAS...74.3735B. doi:10.1073/pnas.74.9.3735. PMC 431708. PMID 198801.
  163. ^ Hayat, M. A. (2007). Cancer imaging: Lung and breast carcinomas. Academic Press. p. 41. ISBN 9780123742124.
  164. ^ Nelson, J. H. (2003). Nuclear magnetic resonance spectroscopy. Prentice Hall. pp. 129–139. ISBN 0-13-033451-0.
  165. ^ 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: 163–195. doi:10.1146/annurev.bb.25.060196.001115. PMC 2899692. PMID 8800468.
  166. ^ Kuethe, Dean O.; Caprihan, Arvind; Fukushima, Eiichi; Waggoner, R. Allen (2005). "Imaging lungs using inert fluorinated gases". Magnetic Resonance in Medicine. 39: 85–88. doi:10.1002/mrm.1910390114.
  167. ^ Shaffer, Thomas H.; Wolfson, Marla R.; Greenspan, Jay S. (1999). "Liquid ventilation: Current status". Pediatrics in Review. 20 (12): e134–e142.
  168. ^ Gabriel, J. L.; Miller, T. F.; Wolfson, M. R. Jr; Shaffer, T. H. (1996). "Quantitative structure-activity relationships of perfluorinated hetro-hydrocarbons as potential respiratory media. Application to oxygen solubility, partition coefficient, viscosity, vapor pressure, and density". 42 (6): 968–973. {{cite journal}}: Cite journal requires |journal= (help); Unknown parameter |lournal= ignored (help)CS1 maint: multiple names: authors list (link)
  169. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.4103/0972-5229.43685, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.4103/0972-5229.43685 instead.
  170. ^ Brown University Division of Biology and Medicine (2006). "History". Retrieved 3 December 2010.
  171. ^ Schimmeyer, S. (2002). "The search for a blood substitute". Illumin. 5 (1). University of Southern Carolina. Retrieved 2 December 2010.
  172. ^ Fred Tasker (2008). "Miami Herald: Artificial blood goes from science fiction to science fact". Miami Herald (at NoBlood.org). Archived from the original on 19 March 2008.
  173. ^ Davis, Nicole (2006). "Better than blood". Popular Science. Retrieved 30 September 2012.
  174. ^ Shaffer, T. H.; Wolfson, M. R.; Clark, L. R. (1992). "State of art review: Liquid ventilation". Pediatric Pulmonology. 14 (102–109). {{cite journal}}: |access-date= requires |url= (help)CS1 maint: multiple names: authors list (link)
  175. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1164/rccm.200508-1196OC, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1164/rccm.200508-1196OC instead.
  176. ^ Kylstra, J. A. (1977). The feasibility of liquid breathing in man. Duke University. Retrieved 5 May 2008.
  177. ^ The Global Oneness Commitment. "Liquid breathing – Space travel". experiencefestival.com. Retrieved 17 May 2008.
  178. ^ Aljean Harmetz (1989). "FILM; 'The Abyss': A foray into deep waters". The New York Times. Retrieved 2 October 2012.
  179. ^ "UN/NA 1045 (United Nations/North America fluorine data sheet)". Cameo Chemicals. National Oceanic and Atmospheric Administration. Retrieved 17 June 2011.
  180. ^ 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.
  181. ^ IPCS (2002). Environmental health criteria 227 (Fluoride). International Programme on Chemical Safety, World Health Organization. p. 100. ISBN 92-4-157227-2.
  182. ^ a b c d Nochimson, Geofrey (2011). "Fluoride toxicity". emedicine. Retrieved 19 May 2011.
  183. ^ Reddy, D. R. (2009). "Neurology of endemic skeletal fluorosis". Neurology India. 57 (1): 7–12. doi:10.4103/0028-3886.48793. PMID 19305069.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  184. ^ 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.
  185. ^ Augenstein, W. L.; et al. (1991). "Fluoride ingestion in children: a review of 87 cases". Pediatrics. 88 (5): 907–912. PMID 1945630. {{cite journal}}: Explicit use of et al. in: |author= (help)
  186. ^ Gessner, Bradford D.; Beller, Michael; Middaugh, John P.; Whitford, Gary M. (1994). "Acute fluoride poisoning from a public water system". New England Journal of Medicine. 330 (2): 95–99. doi:10.1056/NEJM199401133300203. PMID 8259189. {{cite journal}}: |access-date= requires |url= (help)
  187. ^ Sullivan, John Burke; Krieger, Gary R. (2001). Clinical environmental health and toxic exposures. Lippincott Williams & Wilkins. pp. 458–459. ISBN 9780683080278.{{cite book}}: CS1 maint: multiple names: authors list (link)
  188. ^ Blodgett, David W.; Suruda, Anthony J.; Crouch, Barbara Insley (2001). "Fatal unintentional occupational poisonings by hydrofluoric acid in the U.S" (PDF). American Journal of Industrial Medicine. 40 (2): 215–220. doi:10.1002/ajim.1090. PMID 11494350.
  189. ^ Goldfrank's manual of toxicologic emergencies. McGraw-Hill Professional. 2007. p. 1333. ISBN 978-0-07-144310-4. Retrieved 3 May 2011.
  190. ^ a b "Recommended medical treatment for hydrofluoric acid exposure" (PDF). Honeywell Specialty Materials. pp. 2–12. Retrieved 6 May 2009.
  191. ^ 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.
  192. ^ 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): 889–894. doi:10.1097/01.BCR.0000245767.54278.09. PMID 17091088.
  193. ^ 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.
  194. ^ Aucamp, Pieter J.; Björn, Lars Olof. "Questions and answers about the environmental effects of the ozone layer depletion and climate change: 2010 update" (PDF). United Nations Environmental Programme. pp. 4–6, 41, 46–47. Retrieved 6 January 2012.{{cite web}}: CS1 maint: multiple names: authors list (link)
  195. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1038/news.2011.293, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1038/news.2011.293 instead.
  196. ^ Barry, Patrick L.; Phillips, Tony (2006). "Good news and a puzzle". NASA. Retrieved 6 January 2012.
  197. ^ "Science". Ozone Layer Protection. U.S. EPA. Retrieved 6 January 2012.
  198. ^ McCoy, Michael (2007). "Nations speed up HCFC elimination". Chemical & Engineering News. 85 (40): 11. Retrieved 25 December 2012.
  199. ^ "Phaseout of HCFCs (Class II Ozone-Depleting Substances)". U.S. Environmental Protection Agency. 2012. Retrieved 15 August 2012.
  200. ^ "Greenhouse gases: Some definitions". gdrc.org. Retrieved 20 February 2011.
  201. ^ Intergovernmental Panel on Climate Change. "Changes in Atmospheric Constituents and in Radiative Forcing.". Climate change 2007: The physical science basis. Contribution of Working Group I to the fourth assessment report of the Intergovernmental Panel on Climate (PDF). Cambridge University. p. 212. ISBN 978-0-521-70596-7.
  202. ^ a b c 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. Retrieved 3 May 2011.
  203. ^ a b c Betts, Kellyn S. (2007). "Perfluoroalkyl acids: What is the evidence telling us?". Environmental Health Perspectives. 115 (5): A250–A256. doi:10.1289/ehp.115-a250. PMC 1867999. PMID 17520044. Retrieved 5 September 2011.
  204. ^ Calderazzo, Fausto (2010). "Halide-bridged polymers of divalent metals with donor ligands – Structures and properties". Coordination Chemistry Reviews. 254 (5–6): 537–554. doi:10.1016/j.ccr.2009.08.007.
  205. ^ Wiberg, Wiberg & Holleman 2001, p. 422.
  206. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1039/C1RA00804H, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1039/C1RA00804H instead.
  207. ^ a b c d Attention: This template ({{cite pmid}}) is deprecated. To cite the publication identified by PMID 11792193, please use {{cite journal}} with |pmid=11792193 instead.
  208. ^ Smart, Bruce E.; Tatlow, J. C. (1994). Organofluorine chemistry: Principles and commercial application. Springer. p. 515. ISBN 9780306646108. {{cite book}}: Check |isbn= value: checksum (help)CS1 maint: multiple names: authors list (link)
  209. ^ Pauling, Linus A. (1960). The nature of the chemical bond and the structure of molecules and crystals: An introduction to modern structural chemistry. Cornell University Press. pp. 454–464. ISBN 978-0-8014-0333-0.
  210. ^ Atkins, Peter (2008). Chemical principles: The quest for insight. W. H. Freeman & Co. pp. 184–185. ISBN 978-1-4292-0965-6. {{cite book}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  211. ^ Emsley, John (1981). "The hidden strength of hydrogen". New Scientist. 91 (1264): 291–292. Retrieved 25 December 2012.
  212. ^ Greenwood & Earnshaw 1998, pp. 812–816.
  213. ^ Wiberg, Wiberg & Holleman 2001, p. 425.
  214. ^ Clark, Jim. "The acidity of the hydrogen halides". Retrieved 4 September 2011.
  215. ^ a b Chambers, C.; Holliday, A. K. (1975). Modern inorganic chemistry (An intermediate text) (PDF). The Butterworth Group. pp. 328–329.{{cite book}}: CS1 maint: multiple names: authors list (link)
  216. ^ Hannan, Henry J. (1975). Technician's formulation handbook for industrial and household cleaning Products. Lulu.com. p. 31. ISBN 9780615156019.
  217. ^ Hannan, Henry J. (2010). Course in chemistry for IIT-JEE 2011. Tata McGraw Hill Education Private Limited. pp. 15–22. ISBN 9780070703360.
  218. ^ Greenwood & Earnshaw 1998, pp. 816–819.
  219. ^ Greenwood & Earnshaw 1998, p. 819.
  220. ^ Lide 2004, p. 4-76.
  221. ^ Lide 2004, p. 4-71.
  222. ^ Lide 2004, p. 4-92.
  223. ^ Oxtoby, David W.; Gillis, H. Pat; Campion, Alan (2012). Principle of modern chemistry. Cengage learning. p. 693. ISBN 9780840049315.{{cite book}}: CS1 maint: multiple names: authors list (link)
  224. ^ Whitten, Kenneth W.; Davis, Raymond E.; Stanley, George G.; Peck, Larry M. Chemistry (8th ed.). Cengage Learning. p. 351. ISBN 9780495011965.{{cite book}}: CS1 maint: multiple names: authors list (link)
  225. ^ Aigueperse et al. 2005, pp. 25–27.
  226. ^ Arai, Toshihiro (1999). Mesoscopic materials and clusters: Their physical and chemical properties. Springer. p. 267. ISBN 9783540648840.
  227. ^ Remy, Heinrich (1956). Treatise on inorganic chemistry: Introduction and main groups of the periodic table. Elsevier Publishing Company. p. 383.
  228. ^ Wiberg, Wiberg & Holleman 2001, p. 1268.
  229. ^ Emeléus & Sharpe 1983, pp. 256–277.
  230. ^ Walsh, Kenneth A. Beryllium chemistry and processing. ASM International. pp. 99–102, 118–119. ISBN 978-0-87170-721-5.
  231. ^ Mackay, Mackay & Henderson 2002, pp. 243–244.
  232. ^ Hertz, Raymond K. (1987). "General analytical chemistry of beryllium". In Coyle, Francis T. (ed.). Chemical analysis of metals: A symposium. ASTM. pp. 74–75. ISBN 978-0-8031-0942-1.
  233. ^ Wiberg, Wiberg & Holleman 2001, pp. XXXVI–XXXVII.
  234. ^ Wiberg, Wiberg & Holleman 2001, p. XXXVI.
  235. ^ Lide 2004, p. 12-23.
  236. ^ Wiberg, Wiberg & Holleman 2001, p. 1073.
  237. ^ Wiberg, Wiberg & Holleman 2001, p. 1047.
  238. ^ Wiberg, Wiberg & Holleman 2001, p. 1286.
  239. ^ Sobolev, Boris Petrovich (2001). The rare earth trifluorides: Introduction to materials science of multicomponent metal fluoride crystals. Institut d'Estudis Catalans. p. 51. ISBN 84-7283-610-X.
  240. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1063/1.467963, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1063/1.467963 instead.
  241. ^ 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.{{cite book}}: CS1 maint: multiple names: authors list (link)
  242. ^ Lide 2004, pp. 4–60, 4–76, 4–92, 4–96.
  243. ^ Greenwood & Earnshaw 1998, p. 964.
  244. ^ Greenwood & Earnshaw 1998, p. 990.
  245. ^ Becker, S.; Muller, B. G. (1990). "Vanadium tetrafluoride". Angewandte Chemie International Edition (29): 406. {{cite journal}}: More than one of |number= and |issue= specified (help)
  246. ^ "Vanadium tetrafluoride". Retrieved 13 May 2012.
  247. ^ Greenwood & Earnshaw 1998, p. 1120.
  248. ^ a b c Greenwood & Earnshaw 1998, p. 1083.
  249. ^ Brown, David; Canterford, J. H.; Colton, Ray (2009). Halides of the transition elements: Halides of the first row transition metals. Wiley. p. 213. Retrieved 6 June 2012.{{cite book}}: CS1 maint: multiple names: authors list (link)
  250. ^ Yaws & Braker 2001, p. 397.
  251. ^ a b Institut für Chemie (2012). "Prediction of new compounds and new oxidation states". Technische Universität Berlin. Retrieved 24 May 2012.
  252. ^ Breunig, Hans Joachim. "Bismuth compounds". Kirk-Othmer encyclopedia of chemical technology. Vol. 4 (5th ed.). John Wiley & Sons. p. 22. ISBN 978-0-471-48519-3.
  253. ^ Wiberg, Wiberg & Holleman 2001, p. 770.
  254. ^ Suzuki, Hitomi; Matano, Yoshihiro (2001). Organobismuth chemistry. Elsevier. p. 8. ISBN 0-444-20528-4.
  255. ^ Greenwood & Earnshaw 1998, p. 989.
  256. ^ Bartlett, Neil; Lohmann, D. H. (1962). "Dioxygenyl hexafluoroplatinate (V), O2+[PtF6]". Proceedings of the Chemical Society (3). Chemical Society: 115. doi:10.1039/PS9620000097.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  257. ^ Bartlett, Neil (1962). "Xenon hexafluoroplatinate (V) Xe+[PtF6]". Proceedings of the Chemical Society (6). Chemical Society: 218. doi:10.1039/PS9620000197.
  258. ^ Wiberg, Wiberg & Holleman 2001, p. 594.
  259. ^ Vogt, T.; Fitch, A. N.; Cockcroft, J. K. (1994). "Crystal and molecular structures of rhenium heptafluoride". Science. 263 (5151): 1265–1267. Bibcode:1994Sci...263.1265V. doi:10.1126/science.263.5151.1265. PMID 17817431.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  260. ^ Greenwood & Earnshaw 1998, pp. 1061–1062.
  261. ^ Wiberg, Wiberg & Holleman 2001, p. 1420.
  262. ^ Bayerische Julius-Maximilians-Universität Würzburg 2006, p. 93.
  263. ^ a b c Riedel, S.; Kaupp, M. (2009). "The highest oxidation states of the transition metal elements". Coordination Chemistry Reviews. 253: 606–624. doi:10.1016/j.ccr.2008.07.014.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  264. ^ Bayerische Julius-Maximilians-Universität Würzburg 2006, p. 102.
  265. ^ Noury, Stephane; Silvi, Bernard; Gillespie, Ronald J. (2002). "Chemical bonding in hypervalent molecules: Is the octet rule relevant?" (PDF). Inorganic Chemistry. 41: 2164–2172. Retrieved 23 May 2012.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  266. ^ Martinie, R. J.; Bultema, J. J.; van der Wal, M. N.; Burkhart, B. J.; van der Griend, D. A.; de Kock, R. L. (2011). "Bond order and chemical properties of BF, CO, and N2" (PDF). Journal of Chemical Education. 88 (8): 1094–1097. doi:10.1021/ed100758t.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  267. ^ Greenwood & Earnshaw 1998, pp. 198–199.
  268. ^ Ellis, Bryan David (2001). Scientific essentialism. Cambridge University Press. p. 69. ISBN 0521800943.
  269. ^ a b Aigueperse et al. 2005, p. 28.
  270. ^ Aigueperse et al. 2005, p. 30.
  271. ^ a b c Raghavan, P. S. (1998). Concepts and problems in inorganic Chemistry. Discovery Publishing House. pp. 164–165. ISBN 978817141418. {{cite book}}: Check |isbn= value: length (help)
  272. ^ a b Aigueperse et al. 2005, p. 37.
  273. ^ a b Norman, Nicholas C. (1998). Chemistry of arsenic, antimony and bismuth. Springer. p. 97. ISBN 075140389X.
  274. ^ a b c Murthy, C. Parameshwara. University chemistry, Tom 1. New Age International. pp. 180–182, 206–208. ISBN 8122407420.
  275. ^ Christe, K. O.; Wilson, W. W. (1986). "Synthesis and characterization of NF+
    4
    BrF
    4
    and NF+
    4
    BrF
    4
    O
    ". Inorganic Chemistry. 25 (11): 1904–1906. doi:10.1021/ic00231a038.
  276. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1016/S0022-1139(99)00139-6, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1016/S0022-1139(99)00139-6 instead.
  277. ^ Wiberg, Wiberg & Holleman 2001, p. 435.
  278. ^ Wiberg, Wiberg & Holleman 2001, p. 436.
  279. ^ a b Pitzer, Kenneth Sanborn, ed. (1993). Molecular structure and statistical thermodynamics: Selected papers of Kenneth S. Pitzer. Vol. 1. World Scientific. p. 111. ISBN 9810214391.
  280. ^ Gmelin, Leopold. Gmelin handbook of inorganic chemistry: At—Astatine (8th ed.). Springer-Verlag. p. 224. ISBN 9783540935162.
  281. ^ Greenwood & Earnshaw 1998, pp. 828–830.
  282. ^ Patnaik, Pradyot (2007). A comprehensive guide to the hazardous properties of chemical substances. John Wiley & Sons. pp. 478–479. ISBN 9780471714583.
  283. ^ a b Olah, George A. (2005). "Crossing conventional boundaries in half a century of research". Journal of Organic Chemistry. 70 (7): 2413–2429. doi:10.1021/jo040285o. PMID 15787527.
  284. ^ "The Nobel Prize in chemistry 1994". The Nobel Foundation. Retrieved 22 December 2008.
  285. ^ Wiberg, Wiberg & Holleman 2001, pp. 392–393.
  286. ^ Chemical and Engineering News as cited by Michael Barnes. "Neil Bartlett, emeritus professor of chemistry, dies at 75". University of California Newsroom. Retrieved 24 December 2011.
  287. ^ Wiberg, Wiberg & Holleman 2001, p. 438.
  288. ^ Wiberg, Wiberg & Holleman 2001, p. 400.
  289. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1126/science.139.3559.1047, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1126/science.139.3559.1047 instead.
  290. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1021/ic701313h, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1021/ic701313h instead.
  291. ^ Xu, Ruren; Pang, Wenqin; Huo, Qisheng (2011). Main inorganic synthetic chemistry. p. 54. ISBN 9780444535993.{{cite book}}: CS1 maint: multiple names: authors list (link)
  292. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1063/1.1506150, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1063/1.1506150 instead.
  293. ^ Lias, S. G.; Liebman, J. F.; Levin, R. D. (1984). "Evaluated gas phase basicities and proton affinities of molecules; Heats of formation of protonated molecules". Journal of Physical and Chemical Reference Data. 13 (3): 695. Bibcode:1984JPCRD..13..695L. doi:10.1063/1.555719.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  294. ^ Lewars 2008, pp. 70–78.
  295. ^ 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" (PDF). Journal of Physical Chemistry A. 103 (8): 1104–1108. doi:10.1021/jp983665k.
  296. ^ 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.
  297. ^ Wiberg, Wiberg & Holleman 2001, p. 398.
  298. ^ Polysciences, Inc. (2000). "Technical data sheet 320 Ruthenium tetroxide 0.5% aqueous solution" (PDF). p. 1. Retrieved 28 December 2012.
  299. ^ Bayerische Julius-Maximilians-Universität Würzburg 2006, p. 34.
  300. ^ Weldkamp, Achim (1993). "Quantum-mechanical ab initio investigation of the transition-metal compounds OsO4, OsO3F2, OsO2F4, OsOF6, and OsF8". Chemische Berichte. 126 (6): 1325–1330. doi:10.1002/cber.19931260609. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  301. ^ 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.
  302. ^ Wang, Xuefang; Andrews, Lester; Riedel, Sebastian; Kaupp, Martin (2007). "Mercury is a transition metal: The first experimental evidence for HgF4". Angewandte Chemie. 119 (44): 8523–8527. doi:10.1002/ange.200703710.
  303. ^ Lewars 2008, pp. 60–61.
  304. ^ O'Hagan, D. (2008). "Understanding organofluorine chemistry. An introduction to the C–F bond". Chemical Society Reviews. 37 (2): 308–319. doi:10.1039/b711844a. PMID 18197347.
  305. ^ a b c Siegemund et al. 2005, p. 2.
  306. ^ Siegemund et al. 2005, p. 4.
  307. ^ Jassaud et al. 2005, p. 13.
  308. ^ Siegemund et al. 2005, pp. 3–5.
  309. ^ a b Carlson & Scmiegel 2005, p. 3.
  310. ^ Kharitonov, A. P.; Taege, R.; Ferrier, G.; Piven, N. P. (2005). "The kinetics and mechanism of the direct fluorination of polyethylenes". Surface Coatings International Part B: Coatings Transactions. 88. Oil & Colour Chemists' Association: 201–212. ISSN 1356-0751.
  311. ^ a b Siegemund et al. 2005, pp. 7–8.
  312. ^ a b Sukornick, B. (1989). "Potentially acceptable substitutes for the chlorofluorocarbons". International Journal of Thermophysics. 10 (3): 553–561. Bibcode:1989IJT....10..553S. doi:10.1007/BF00507978.
  313. ^ a b Siegemund et al. 2005, p. 18.
  314. ^ Urben, Peter (2006). Bretherick's handbook of reactive chemical hazards (7th ed.). Surendra Kumar. p. 269. ISBN 9780123725639.
  315. ^ Siegemund et al. 2005, p. 5.
  316. ^ Mendicino, L.; Dielectric Science and Technology Division (Electrochemical Society) (1999). Environmental issues in the electronics and semiconductor industries. The Electrochemical Society. p. 116. ISBN 9781566772303.{{cite book}}: CS1 maint: multiple names: authors list (link)
  317. ^ a b Knepper, Thomas P.; Lange, Frank T. (2011). Springer. p. 27. ISBN 9783642218712. {{cite book}}: Missing or empty |title= (help); Unknown parameter |tit;e= ignored (help)CS1 maint: multiple names: authors list (link)
  318. ^ Siegemund et al. 2005, p. 29.
  319. ^ Introductory Chemistry Online (2011). "Acids and bases". Retrieved 18 August 2012.
  320. ^ a b Siegemund et al. 2005, p. 32.
  321. ^ University of California, Berkeley. "Research into gecko adhesion". Archived from the original on 14 October 2007. Retrieved 29 April 2011.
  322. ^ Carlson & Scmiegel 2005, pp. 3–4.
  323. ^ a b Carlson & Scmiegel 2005, p. 4.
  324. ^ a b c d Carlson & Scmiegel 2005, pp. 26–27.
  325. ^ a b Mittemeijer, Eric J. (2011). Fundamentals of materials science: The microstructure–property relationship using metals as model systems. Springer. p. 515. ISBN 9783642104992.
  326. ^ Rhoades, David Walter; The University of Southern Mississippi (2008). Broadband dielectric spectroscopy studies of Nafion. ProQuest. p. 2. ISBN 9780549785408.{{cite book}}: CS1 maint: multiple names: authors list (link)

Bibliography

Template:Link GA Template:Link GA Template:Link FA