Phosphorus

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Nergaal (talk | contribs) at 02:35, 12 October 2008 (→‎Chemical bonding). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Phosphorus, 15P
Forms of phosphorus
Waxy white
Light red
Dark red and violet
Black
Phosphorus
Pronunciation/ˈfɒsfərəs/ (FOS-fər-əs)
Allotropeswhite, red, violet, black and others (see Allotropes of phosphorus)
Appearancewhite, red and violet are waxy, black is metallic-looking
Standard atomic weight Ar°(P)
Abundance
in the Earth's crust5.2 (silicon = 100)
Phosphorus 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
N

P

As
siliconphosphorussulfur
Atomic number (Z)15
Groupgroup 15 (pnictogens)
Periodperiod 3
Block  p-block
Electron configuration[Ne] 3s2 3p3
Electrons per shell2, 8, 5
Physical properties
Phase at STPsolid
Melting pointwhite: 317.3 K ​(44.15 °C, ​111.5 °F)
red: ∼860 K (∼590 °C, ∼1090 °F)[3]
Boiling pointwhite: 553.7 K ​(280.5 °C, ​536.9 °F)
Sublimation pointred: ≈689.2–863 K ​(≈416–590 °C, ​≈780.8–1094 °F)
violet: 893 K (620 °C, 1148 °F)
Density (near r.t.)white: 1.823 g/cm3
red: ≈2.2–2.34 g/cm3
violet: 2.36 g/cm3
black: 2.69 g/cm3
Heat of fusionwhite: 0.66 kJ/mol
Heat of vaporizationwhite: 51.9 kJ/mol
Molar heat capacitywhite: 23.824 J/(mol·K)
Vapor pressure (white)
P (Pa) 1 10 100 1 k 10 k 100 k
at T (K) 279 307 342 388 453 549
vapor pressure (red)
P (Pa) 1 10 100 1 k 10 k 100 k
at T (K) 455 489 529 576 635 704
Atomic properties
Oxidation states−3, −2, −1, 0,[4] +1,[5] +2, +3, +4, +5 (a mildly acidic oxide)
ElectronegativityPauling scale: 2.19
Ionization energies
  • 1st: 1011.8 kJ/mol
  • 2nd: 1907 kJ/mol
  • 3rd: 2914.1 kJ/mol
  • (more)
Covalent radius107±3 pm
Van der Waals radius180 pm
Color lines in a spectral range
Spectral lines of phosphorus
Other properties
Natural occurrenceprimordial
Crystal structureα-white: ​body-centered cubic (bcc) (cI232)
Lattice constant
Body-centred cubic crystal structure for α-white: phosphorus
a = 1.869 nm (at 20 °C)[6]
Crystal structureblack: ​orthorhombic (oS8)
Lattice constants
Orthorhombic crystal structure for black: phosphorus
a = 0.33137 nm
b = 1.0477 nm
c = 0.43755 nm (at 20 °C)[6]
Thermal conductivitywhite: 0.236 W/(m⋅K)
black: 12.1 W/(m⋅K)
Magnetic orderingwhite, red, violet, black: diamagnetic[7]
Molar magnetic susceptibility−20.8×10−6 cm3/mol (293 K)[8]
Bulk moduluswhite: 5 GPa
red: 11 GPa
CAS Number7723-14-0 (red)
12185-10-3 (white)
History
DiscoveryHennig Brand (1669)
Recognised as an element byAntoine Lavoisier[9] (1777)
Isotopes of phosphorus
Main isotopes Decay
abun­dance half-life (t1/2) mode pro­duct
31P 100% stable
32P trace 14.269 d β 32S
33P trace 25.35 d β 33S
 Category: Phosphorus
| references

Phosphorus, (/ˈfɒsfərəs/), is the chemical element that has the symbol P and atomic number 15. The name comes from the [φώς] Error: {{Lang-xx}}: text has italic markup (help) (meaning "light") and φόρος (meaning "bearer"). A multivalent nonmetal of the nitrogen group, phosphorus is commonly found in inorganic phosphate rocks.

Due to its high reactivity, phosphorus is never found as a free element in nature on Earth. One form of phosphorus (white phosphorus) emits a faint glow upon exposure to oxygen — hence its Greek derivation, Φωσφόρος meaning "light-bearer" (Latin Lucifer), the planet Venus as "Morning Star".

Phosphorus is a component of DNA and RNA, as well as ATP, and is an essential element for all living cells. The most important commercial use of phosphorus-based chemicals is the production of fertilizers.

Phosphorus compounds are also widely used in explosives, nerve agents, friction matches, fireworks, pesticides, toothpaste and detergents.

Characteristics

Allotropes

Phosphorus is an excellent example of an element that exhibits allotropy, as its various allotropes have strikingly different properties.

The two most common allotropes are white phosphorus and red phosphorus. A third form, scarlet phosphorus, is obtained by allowing a solution of white phosphorus in carbon disulfide to evaporate in sunlight. A fourth allotrope, black phosphorus, is obtained by heating white phosphorus under very high pressures (12,000 atmospheres). In appearance, properties and structure it is very like graphite, being black and flaky, a conductor of electricity and has puckered sheets of linked atoms. Another allotrope is diphosphorus - which is highly reactive.

White phosphorus (P
4
) exists as individual molecules made up of four atoms in a tetrahedral arrangement, resulting in very high ring strain and instability. It contains 6 single bonds.

White phosphorus is a white, waxy transparent solid. This allotrope is thermodynamically unstable at normal condition and will gradually change to red phosphorus. This transformation, which is accelerated by light and heat, makes white phosphorus almost always contain some red phosphorus and therefore appear yellow. For this reason, it is also called yellow phosphorus. It glows greenish in the dark (when exposed to oxygen), is highly flammable and pyrophoric (self-igniting) upon contact with air as well as toxic (causing severe liver damage on ingestion). The incendiary bomb Napalm relies, among others, on this principle to spontaneously ignite. The odour of combustion of this form has a characteristic garlic smell, and samples are commonly coated with white "(di)phosphorus pentoxide", which consists of P4O10 tetrahedra with oxygen inserted between the phosphorus atoms and at their vertices. White phosphorus is insoluble in water but soluble in carbon disulfide.

The white allotrope can be produced using several different methods. In one process, calcium phosphate, which is derived from phosphate rock, is heated in an electric or fuel-fired furnace in the presence of carbon and silica[10]. Elemental phosphorus is then liberated as a vapour and can be collected under phosphoric acid. This process is similar to the first synthesis of phosphorus from calcium phosphate in urine.

Red phosphorus may be formed by heating white phosphorus to 250°C (482°F) or by exposing white phosphorus to sunlight. Phosphorus after this treatment exists as an amorphous network of atoms which reduces strain and gives greater stability; further heating results in the red phosphorus becoming crystalline. Red phosphorus does not catch fire in air at temperatures below 240°C, whereas white phosphorus ignites at about 30°C.

In 1865 Hittorf discovered that when phosphorus was recrystallized from molten lead, a red/purple form is obtained. This purple form is sometimes known as "Hittorf's phosphorus." In addition, a fibrous form exists with similar phosphorus cages. Below is shown a chain of phosphorus atoms which exhibits both the purple and fibrous forms.

One of the forms of red/black phosphorus is a cubic solid.[11]

Black phosphorus has an orthorhombic structure (Cmca) and is the least reactive allotrope. It consists of many six-membered rings which are interlinked. Each atom is bonded to three other atoms.[12][13] A recent synthesis of black phosphorus using metal salts as catalysts has been reported.[14]

The diphosphorus allotrope (P2) can be obtained normally only under extreme conditions (for example, from P4 at 1100 kelvin). Nevertheless, some advancements were obtained in generating the diatomic molecule in homogeneous solution, under normal conditions with the use by some transitional metal complexes (based on, for example, tungsten and niobium).[15]

Glow

The chemical element phosphorus (Greek. phosphoros, meaning "light bearer") was discovered by German alchemist Hennig Brand in 1669. Working in Hamburg, Brand attempted to distill some kind of "life essence" from his urine, and in the process produced a white material that glowed in the dark. However, phosphorus itself is not a phosphor; it is highly reactive and gives-off a faint chemiluminescent glow upon uniting with oxygen. The glow observed by Brand was actually caused by the very slow burning of the phosphorus, but as he saw no flame nor felt any heat he did not recognize it as burning.

The mechanism for that glow was not fully described until 1974.[16] It was known from early times that the glow would persist for a time in a stoppered jar but then cease. Robert Boyle in the 1680s ascribed it to "debilitation" of the air; in fact, it is oxygen being consumed. By the 18th century, it was known that in pure oxygen phosphorus does not glow at all;[17] there is only a range of partial pressure at which it does. Heat can be applied to drive the reaction at higher pressures.[18]

In 1974, the glow was explained by R. J. van Zee and A. U. Khan.[16] A reaction with oxygen takes place at the surface of the solid (or liquid) phosphorus, forming the short-lived molecules HPO and P2O2 that both emit visible light. The reaction is slow and only very little of the intermediates is required to produce the luminescence, hence the extended time the glow continues in a stoppered jar.

Although the term phosphorescence is derived from phosphorus, the reaction which gives phosphorus its glow is properly called luminescence (glowing by its own reaction, in this case chemoluminescence), not phosphorescence (re-emitting light that previously fell on it).

Isotopes

Although twenty three isotopes of phosphorus are known [19] (all posibilities from 24P up to 46P), only 31P, with spin 1/2, is stable and is therefore present at 100% abundance. The half-integer spin and high abundance of 31P make it useful for nuclear magnetic resonance studies of biomolecules, particularly DNA.

Two radioactive isotopes of phosphorous have half-lives which make them useful for scientific experiments. 32P has a half-life of 14.262 days and 33P has a half-life of 25.34 days. Biomolecules can be "tagged" with a radio isotope to allowed for the study of very dilute samples.

Radioactive isotopes of phosphorus include

  • 32P; a beta-emitter (1.71 MeV) with a half-life of 14.3 days which is used routinely in life-science laboratories, primarily to produce radiolabeled DNA and RNA probes, e.g. for use in Northern blots or Southern blots. Because the high energy beta particles produced penetrate skin and corneas, and because any 32P ingested, inhaled, or absorbed is readily incorporated into bone and nucleic acids, Occupational Safety and Health Administration in the Unites States, and similar institutions in other developed countries require that a lab coat, disposable gloves, and safety glasses or goggles be worn when working with 32P, and that working directly over an open container be avoided in order to protect the eyes. Monitoring personal, clothing, and surface contamination is also required. In addition, due to the high energy of the beta particles, shielding this radiation with the normally used dense materials (e.g. lead), gives rise to secondary emission of X-rays via a process known as Bremsstrahlung, meaning braking radiation. Therefore shielding must be accomplished with low density materials, e.g. Plexiglas, Lucite, plastic, wood, or water.[20]
  • 33P; a beta-emitter (0.25 MeV) with a half-life of 25.4 days. It is used in life-science laboratories in applications in which lower energy beta emissions are advantageous such as DNA sequencing.

Occurrence

See also Phosphate minerals.

Due to its reactivity with air and many other oxygen-containing substances, phosphorus is not found free in nature but it is widely distributed in many different minerals.

Phosphate rock, which is partially made of apatite (an impure tri-calcium phosphate mineral), is an important commercial source of this element. About 50 per cent of the global phosphorus reserves are in the Arab nations.[21] Large deposits of apatite are located in China, Russia, Morocco, Florida, Idaho, Tennessee, Utah, and elsewhere. Albright and Wilson in the United Kingdom and their Niagara Falls plant, for instance, were using phosphate rock in the 1890s and 1900s from Connetable, Tennessee and Florida; by 1950 they were using phosphate rock mainly from Tennessee and North Africa[10]. In the early 1990s Albright and Wilson's purified wet phosphoric acid business was being affected by phosphate rock sales by China and the entry of their long standing Moroccan phosphate suppliers into the purified wet phosphoric acid business.[22]

In 2007, at the current rate of consumption, the supply of phosphorus was estimated to run out in 345 years.[23] However, scientists are now claiming that a "Peak Phosphorus" will occur in 30 years and that "At current rates, reserves will be depleted in the next 50 to 100 years."[24]

Production

Calcium phosphate (phosphate rock), mostly mined in Florida and North Africa, can be heated to 1200-1500 Celsius with sand, which is mostly SiO2, and coke (impure carbon) to produce vaporized tetraphosphorus, P4, (mp. 44.2 C) which is subsequently condensed into a white power under water to prevent oxidation. Even under water, white phosphorus is slowly converted to the more stable red phosphorus allotrope (mp. 597C). Both the white and red allotropes of phosphorus are insoluble in water.

Chemistry

See also Phosphorus compounds

Chemical bonding

Because phosphorus is just below nitrogen in the periodic table, the two elements share many of their bonding characteristics. For instance, phosphine, PH3, is an analogue of ammonia, NH3. Phosphorus, like nitrogen, is trivalent in this molecule,

This is the pre-quantum mechanical Lewis structure, which although somewhat of a simplification from a quantum chemical point of view,[25] illustrates some of the distinguishing characteristics. The unpaired electrons in the three 3p orbitals bind with those in the hydrogen 1s orbitals to form electron pairs of opposite spin. The electron pair in the phosphorus 3s orbital shows up as a lone pair in the Lewis structure.

The phosphorus cation is very similar to the nitrogen cation. In the same way that nitrogen forms the tetravalent ammonium ion, phosphorus can form the tetravalent phosphonium ion, and form salts such as phosphonium iodide [PH4]+[I].

Like other elements in the third or lower rows of the periodic table, phosphorus atoms can expand their valence to make penta- and hexavalent compounds. The phosphorus chloride molecule is an example. When the phosphorus ligands are not identical, the more electronegative ligands are located in the apical positions and the least electronegative ligands are located in the axial positons.

With strongly electronegative ions, in particular fluorine, hexavalency as in PF6 occurs as well. This octahedral ion is isoelectronic with SF6. In the bonding the six octahedral sp3d2 hybrid atomic orbitals play an important role.

Before extensive computer calculations were feasible, it was generally assumed that the nearby d orbitals in the n = 3 shell were the obvious cause of the difference in binding between nitrogen and phosphorus. However, in the early eighties the German theoretical chemist Werner Kutzelnigg[26] found from an analysis of computer calculations that the difference in binding is more likely due to differences in character between the valence 2p and valence 3p orbitals of nitrogen and phosphorus, respectively. The 2s and 2p orbitals of first row atoms are localized in roughly the same region of space, while the 3p orbitals of phosphorus are much more extended in space. The violation of the octet rule observed in compounds of phosphorus is then due to the size of the P-atom and the corresponding reduction of steric hindrance between its ligands. In modern theoretical chemistry, Kutzelnigg's analysis is generally accepted.

The simple Lewis structure for the trigonal bipyramidal PCl5 molecule contains five covalent bonds, implying a hypervalent molecule with ten valence electrons contrary to the octet rule.

An alternate description of the bonding, however, respects the octet rule by using 3-center-4-electron (3c-4e) bonds. In this model the octet on the P atom corresponds to six electrons which form three Lewis (2c-2e) bonds to the three equatorial Cl atoms, plus the two electrons in the 3-centre Cl-P-Cl bonding molecular orbital for the two axial Cl electrons. The two electrons in the corresponding nonbonding molecular orbital are not included because this orbital is localized on the two Cl atoms and does not contribute to the electron density on P.

However, it should always be remembered that the octet rule is not some universal rule of chemical bonding, and while many compounds obey it, there are many elements (the majority, in fact) to which it just does not apply.

Phosphine, diphosphine and phosphonium salts

Phosphine (PH3) and arsine (AsH3) are structural analogs with ammonia (NH3) and form pyramidal structures with the phosphorus or arsenic atom in the center bound to three hydrogen atoms and one lone electron pair. Both are colorless, ill-smelling, toxic compounds. Phosphine is produced in a manner similar to the production of ammonia. Hydrolysis of calcium phosphide, Ca3P2, or calcium nitride, Ca3N2 produce phosphine or ammonia, respectively. Unlike ammonia, phosphine is unstable and it reacts instantly with air giving off phosphoric acid clouds. Arsenic is even less stable. Although phosphine is less basic than ammonia, it can form some phosphonium salts (like PH4I), analogs of ammonium salts, but these salts immediately decompose in water and do not yield phosphonium (PH4+) ions. Diphosphine (P2H4 or H2P-PH2) is an analog of hydrazine (N2H4) that is a colorless liquid which spontaneously ignites in air and can disproportionate into phosphine and complex hydrides.

Applications

Biological role

Phosphorus is a key element in all known forms of life. Inorganic phosphorus in the form of the phosphate PO43- plays a major role in biological molecules such as DNA and RNA where it forms part of the structural framework of these molecules. Living cells also use phosphate to transport cellular energy via adenosine triphosphate (ATP). Nearly every cellular process that uses energy obtains it in the form of ATP. ATP is also important for phosphorylation, a key regulatory event in cells. Phospholipids are the main structural components of all cellular membranes. Calcium phosphate salts assist in stiffening bones.

An average adult human contains a little less than 1 kg of phosphorus, about 85% of which is present in bones and teeth in the form of apatite, and the remainder inside cells in soft tissues. A well-fed adult in the industrialized world consumes and excretes about 1-3 g of phosphorus per day in the form of phosphate. Only about 0.1% of body phosphate circulates in the blood, but this amount reflects the amount of phosphate available to soft tissue cells.

In medicine, low phosphate syndromes are caused by malnutrition, by failure to absorb phosphate, and by metabolic syndromes which draw phosphate from the blood or pass too much of it into the urine. All are characterized by hypophosphatemia (see article for medical details). Symptoms of low phosphate include muscle and neurological dysfunction, and disruption of muscle and blood cells due to lack of ATP. Too much phosphate can lead to diarrhea and calcification (hardening) of organs and soft tissue, and can interfere with the body's ability to use iron, calcium, magnesium, and zinc.[27]

Phosphorus is an essential macromineral for plants, which is studied extensively in edaphology in order to understand plant uptake from soil systems. In ecological terms, phosphorus is often a limiting factor in many environments; i.e. the availability of phosphorus governs the rate of growth of many organisms. In ecosystems an excess of phosphorus can be problematic, especially in aquatic systems, see eutrophication and algal blooms.

History

Phosphorus (Greek phosphoros was the ancient name for the planet Venus, but in Greek mythology, Hesperus and Eosphorus could be confused with Phosphorus) was discovered by German alchemist Hennig Brand in 1669 through a preparation from urine, which contains considerable quantities of dissolved phosphates from normal metabolism. Working in Hamburg, Brand attempted to create the infamous Philosopher's stone through the distillation of some salts by evaporating urine, and in the process produced a white material that glowed in the dark and burned brilliantly. Since that time, phosphorescence has been used to describe substances that shine in the dark without burning.

Phosphorus was recognized as a chemical element at the emergence of the atomic theory that gradually occurred in the late part of the 18th century and the early 19th century, and was formulated by John Dalton.

Phosphorus was discovered by the Hamburg alchemist Hennig Brand in 1669. He made it out of human urine. He let urine stand for days until it gave off a terrible smell. Then he boiled it down to a paste, heated this paste to a high temperature, and led the vapors through water where he hoped they would condense to gold. Instead, he obtained a white, waxy substance that glowed in the dark. Brand had discovered phosphorus, the first element discovered since antiquity. The word phosphorus comes from the Greek (φως = light, φορέω = carry) and means light carrier. We now know that Brand produced ammonium sodium hydrogenphosphate (NH4)NaHPO4.

Brand at first tried to keep the method secret,[28] but later sold the recipe for 200 thaler to somebody from Dresden, Krafft, who could now make it as well, and toured much of Europe with it, including England, where he met with Robert Boyle. The secret that it was made from urine leaked out and first Johann Kunckel (1630-1703) in Sweden (1678) and later Boyle in London (1680) also managed to make phosphorus. Boyle states that Krafft gave him no no information as to the preparation of phosphorus other than that it was derived from "somewhat that belonged to the body of man". This gave Boyle a valuable clue, however, so that he, too, managed to make phosphorus. Later he improved Brand's process by using sand in the reaction (still using urine as base material),

4NaPO3 + 2SiO2 + 10C → 2Na2SiO3 + 10CO + P4

Robert Boyle was the first to use phosphorus to ignite sulphur-tipped wooden splints, forerunners of our modern matches, in 1680.

In 1769 Johan Gottlieb Gahn and Carl Wilhelm Scheele showed that calcium phosphate (Ca3(PO4)2) is found in bones and they obtained phosphorus from bone ash. Antoine Lavoisier recognized phosphorus as an element in 1777. Bone ash was the major source of phosphorus until the 1840s. Phosphate rock, a mineral containing calcium phosphate, was first used in 1850 and following the introduction of the electric arc furnace in 1890 this became the only source of phosphorus. Phosphorus, phosphates and phosphoric acid are still obtained from phosphate rock. Phosphate rock is a major feedstock in the fertilizer industry.

Phosphorus was first made commercially, for the match industry, in the 19th century, by distilling off phosphorus vapor from precipitated phosphates heated in a retort.[10] The precipitated phosphates were made from ground-up bones that had been de-greased and treated with strong acids.[10] This process became obsolete in the late 1890s when the electric arc furnace was adapted to reduce phosphate rock.[10]

Early matches used white phosphorus in their composition, which was dangerous due to its toxicity. Murders, suicides and accidental poisonings resulted from its use. (An apocryphal tale tells of a woman attempting to murder her husband with white phosphorus in his food, which was detected by the stew giving off luminous steam).[16] In addition, exposure to the vapours gave match workers a necrosis of the bones of the jaw, the infamous "phossy jaw." When a safe process for manufacturing red phosphorus was discovered, with its far lower flammability and toxicity, laws were enacted, under the Berne Convention (1906), requiring its adoption as a safer alternative for match manufacture.

The electric furnace method allowed production to increase to the point where phosphorus could be used in weapons of war.[16][10] In World War I it was used in incendiaries, smoke screens and tracer bullets.[10] A special incendiary bullet was developed to shoot at hydrogen-filled Zeppelins over Britain (hydrogen being highly inflammable if it can be ignited).[10] During World War II, Molotov cocktails of benzene and phosphorus were distributed in Britain to specially selected civilians within the British resistance operation, for defence; and phosphorus incendiary bombs were used in war on a large scale. Burning phosphorus is difficult to extinguish and if it splashes onto human skin it has horrific effects (see precautions below). People covered in it have been known to commit suicide due to the torment.

Today phosphorus production is larger than ever. It is used as a precursor for various chemicals,[29] in particular the herbicide glyphosate sold under the brand name Roundup. Production of white phosphorus takes place at large facilities and it is transported heated in liquid form. Some major accidents have occurred during transportation, train derailments at Brownston, Nebraska and Miamisburg, Ohio led to large fires. The worst accident in recent times was an environmental one in 1968 when phosphorus spilled into the sea from a plant at Placentia Bay, Newfoundland.

Spelling and etymology

According to the Oxford English Dictionary the correct spelling of the element is phosphorus. The word phosphorous is the adjectival form of the P3+ valency: so, just as sulfur forms sulfurous and sulfuric compounds, phosphorus forms phosphorous compounds (see e.g. phosphorous acid) and P5+ valency phosphoric compounds (see e.g. Phosphoric acids and phosphates).

Precautions

Organic compounds of phosphorus form a wide class of materials, some of which are extremely toxic. Fluorophosphate esters are among the most potent neurotoxins known. A wide range of organophosphorus compounds are used for their toxicity to certain organisms as pesticides (herbicides, insecticides, fungicides, etc.) and weaponized as nerve agents. Most inorganic phosphates are relatively nontoxic and essential nutrients. For environmentally adverse effects of phosphates see eutrophication and algal blooms.

The white phosphorus allotrope should be kept under water at all times as it presents a significant fire hazard due to its extreme reactivity with atmospheric oxygen, and it should only be manipulated with forceps since contact with skin can cause severe burns. Chronic white phosphorus poisoning leads to necrosis of the jaw called "phossy jaw". Ingestion of white phosphorus may cause a medical condition known as "Smoking Stool Syndrome". [30]

When the white form is exposed to sunlight or when it is heated in its own vapour to 250°C, it is transmuted to the red form, which does not phosphoresce in air. The red allotrope does not spontaneously ignite in air and is not as dangerous as the white form. Nevertheless, it should be handled with care because it reverts to white phosphorus in some temperature ranges and it also emits highly toxic fumes that consist of phosphorus oxides when it is heated.

Upon exposure to elemental phosphorus, in the past it was suggested to wash the affected area with 2% copper sulfate solution to form harmless compounds that can be washed away. According to the recent US Navy's Treatment of Chemical Agent Casualties and Conventional Military Chemical Injuries: FM8-285: Part 2 Conventional Military Chemical Injuries, "Cupric (copper(II)) sulfate has been used by U.S. personnel in the past and is still being used by some nations. However, copper sulfate is toxic and its use will be discontinued. Copper sulfate may produce kidney and cerebral toxicity as well as intravascular hemolysis."[31]

Phosphorus explosion

The manual suggests instead "a bicarbonate solution to neutralize phosphoric acid, which will then allow removal of visible WP. Particles often can be located by their emission of smoke when air strikes them, or by their phosphorescence in the dark. In dark surroundings, fragments are seen as luminescent spots." Then, "Promptly debride the burn if the patient's condition will permit removal of bits of WP which might be absorbed later and possibly produce systemic poisoning. DO NOT apply oily-based ointments until it is certain that all WP has been removed. Following complete removal of the particles, treat the lesions as thermal burns." As white phosphorus readily mixes with oils, any oily substances or ointments are not recommended until the area is thoroughly cleaned and all white phosphorus removed.

Further warnings of toxic effects and recommendations for treatment can be found in the Emergency War Surgery NATO Handbook: Part I: Types of Wounds and Injuries: Chapter III: Burn Injury: Chemical Burns And White Phosphorus injury.[32]

US DEA List I status

Phosphorus can reduce elemental iodine to hydroiodic acid, which is a reagent effective for reducing ephedrine or pseudoephedrine to methamphetamine.[33] For this reason, two allotropes of elemental phosphorus—red phosphorus and white phosphorus—were designated by the United States Drug Enforcement Administration as List I precursor chemicals under 21 CFR 1310.02 effective November 17, 2001.[34] As a result, in the United States, handlers of red phosphorus or white phosphorus are subject to stringent regulatory controls pursuant to the Controlled Substances Act in order to reduce diversion of these substances for use in clandestine production of controlled substances.[34][35][36]

References

  1. ^ "Standard Atomic Weights: Phosphorus". CIAAW. 2013.
  2. ^ Prohaska, Thomas; Irrgeher, Johanna; Benefield, Jacqueline; Böhlke, John K.; Chesson, Lesley A.; Coplen, Tyler B.; Ding, Tiping; Dunn, Philip J. H.; Gröning, Manfred; Holden, Norman E.; Meijer, Harro A. J. (2022-05-04). "Standard atomic weights of the elements 2021 (IUPAC Technical Report)". Pure and Applied Chemistry. doi:10.1515/pac-2019-0603. ISSN 1365-3075.
  3. ^ Phosphorus at the Encyclopædia Britannica
  4. ^ Wang, Yuzhong; Xie, Yaoming; Wei, Pingrong; King, R. Bruce; Schaefer, Iii; Schleyer, Paul v. R.; Robinson, Gregory H. (2008). "Carbene-Stabilized Diphosphorus". Journal of the American Chemical Society. 130 (45): 14970–1. doi:10.1021/ja807828t. PMID 18937460.
  5. ^ Ellis, Bobby D.; MacDonald, Charles L. B. (2006). "Phosphorus(I) Iodide: A Versatile Metathesis Reagent for the Synthesis of Low Oxidation State Phosphorus Compounds". Inorganic Chemistry. 45 (17): 6864–74. doi:10.1021/ic060186o. PMID 16903744.
  6. ^ a b Arblaster, John W. (2018). Selected Values of the Crystallographic Properties of Elements. Materials Park, Ohio: ASM International. ISBN 978-1-62708-155-9.
  7. ^ Lide, D. R., ed. (2005). "Magnetic susceptibility of the elements and inorganic compounds". CRC Handbook of Chemistry and Physics (PDF) (86th ed.). Boca Raton (FL): CRC Press. ISBN 0-8493-0486-5.
  8. ^ Weast, Robert (1984). CRC, Handbook of Chemistry and Physics. Boca Raton, Florida: Chemical Rubber Company Publishing. pp. E110. ISBN 0-8493-0464-4.
  9. ^ cf. "Memoir on Combustion in General" Mémoires de l'Académie Royale des Sciences 1777, 592–600. from Henry Marshall Leicester and Herbert S. Klickstein, A Source Book in Chemistry 1400–1900 (New York: McGraw Hill, 1952)
  10. ^ a b c d e f g h i j k l m n Threlfall, R.E., (1951). 100 years of Phosphorus Making: 1851 - 1951. Oldbury: Albright and Wilson Ltd
  11. ^ Ahuja, R. (2003). "Calculated high pressure crystal structure transformations for phosphorus". Physica Status Solidi Section B. 235 (2): 282–287. doi:10.1002/pssb.200301569.
  12. ^ A. Brown, S. Runquist (1965). "Refinement of the crystal structure of black phosphorus". Acta Crystallogr. 19: 684. doi:10.1107/S0365110X65004140.
  13. ^ Cartz, L.;Srinivasa, S.R.;Riedner, R.J.;Jorgensen, J.D.;Worlton, T.G. (1979). "Effect of pressure on bonding in black phosphorus". Journal of Chemical Physics. 71: 1718–1721. doi:10.1063/1.438523.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  14. ^ Stefan Lange, Peer Schmidt, and Tom Nilges (2007). "Au3SnP7@Black Phosphorus: An Easy Access to Black Phosphorus". Inorg. Chem. 46: 4028. doi:10.1021/ic062192q.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  15. ^ Piro, N. A. (2006). "Triple-Bond Reactivity of Diphosphorus Molecules". Science. 313 (5791): 1276. doi:10.1126/science.1129630. PMID 16946068.
  16. ^ a b c d Emsley, John (2000). The Shocking History of Phosphorus. London: Macmillan. ISBN 0-330-39005-8
  17. ^ Nobel Prize in Chemistry 1956 - Presentation Speech, by Professor A. Ölander (committee member)
  18. ^ Phosphorus Topics page, at Lateral Science
  19. ^ The Berkeley Laboratory Isotopes Project [1]
  20. ^ http://www.oseh.umich.edu/TrainP32.pdf
  21. ^ http://www.anba.com.br/ingles/noticia.php?id=17288
  22. ^ Podger, Hugh, (2002). Albright & Wilson: The Last 50 Years. Studley: Brewin Books. ISBN 1-85858-223-7
  23. ^ "How Long Will it Last?". New Scientist. 194 (2605): 38–39. May 26, 2007. ISSN 0262-4079. {{cite journal}}: Check date values in: |date= (help)
  24. ^ Leo Lewis (2008-06-23). "Scientists warn of lack of vital phosphorus as biofuels raise demand". The Times.
  25. ^ In quantum chemical valence bond theory one usually works with mixtures of s and p atomic orbitals, so-called hybrids
  26. ^ W. Kutzelnigg, Chemical Bonding in Higher Main Group Elements, Angewandte Chemie Int. (English) Ed., vol. 23, pp. 272-295 (1984)
  27. ^ Anderson JJB. Calcium, phosphorus, and human bone development. J Nutr. 1996; 126: 1153.
  28. ^ J. M. Stillman, The Story of Alchemy and Early Chemistry, Dover, New York (1960), pp. 418-419
  29. ^ Aall C. H. (1952). "The American Phosphorus Industry". Industrial & Engineering Chemistry. 44 (7): 1520–1525. doi:10.1021/ie50511a018.
  30. ^ emedicine.com CBRNE - Incendiary Agents, White Phosphorus (Smoking Stool Syndrome)
  31. ^ US Navy's Treatment of Chemical Agent Casualties and Conventional Military Chemical Injuries: FM8-285: Part 2 Conventional Military Chemical Injuries
  32. ^ Emergency War Surgery NATO Handbook: Part I: Types of Wounds and Injuries: Chapter III: Burn Injury: Chemical Burns And White Phosphorus injury.
  33. ^ Skinner, H.F. (1990). "Methamphetamine synthesis via hydriodic acid/red phosphorus reduction of ephedrine". Forensic Science International. 48 (2): 123–134. doi:10.1016/0379-0738(90)90104-7.
  34. ^ a b 66 FR 52670—52675. 17 October 2001.
  35. ^ 21 CFR 1309
  36. ^ 21 USC, Chapter 13 (Controlled Substances Act)

External links