Rhenium

Last updated
Rhenium, 75Re
Rhenium single crystal bar and 1cm3 cube.jpg
Rhenium
Pronunciation /ˈrniəm/ (REE-nee-əm)
Appearancesilvery-grayish
Standard atomic weight Ar°(Re)
Rhenium 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
Tc

Re

Bh
tungstenrheniumosmium
Atomic number (Z)75
Group group 7
Period period 6
Block   d-block
Electron configuration [ Xe ] 4f14 5d5 6s2
Electrons per shell2, 8, 18, 32, 13, 2
Physical properties
Phase at  STP solid
Melting point 3459  K (3186 °C,5767 °F)
Boiling point 5903 K(5630 °C,10,170 °F)
Density (near  r.t.)21.02 g/cm3
when liquid (at  m.p.)18.9 g/cm3
Heat of fusion 60.43  kJ/mol
Heat of vaporization 704 kJ/mol
Molar heat capacity 25.48 J/(mol·K)
Vapor pressure
P (Pa)1101001 k10 k100 k
at T (K)330336144009450051275954
Atomic properties
Oxidation states −3, −1, 0, +1, +2, +3, +4, +5, +6, +7 (a mildly acidic oxide)
Electronegativity Pauling scale: 1.9
Ionization energies
  • 1st: 760 kJ/mol
  • 2nd: 1260 kJ/mol
  • 3rd: 2510 kJ/mol
  • (more)
Atomic radius empirical:137  pm
Covalent radius 151±7 pm
Rhenium spectrum visible.png
Spectral lines of rhenium
Other properties
Natural occurrence primordial
Crystal structure hexagonal close-packed (hcp)
Hexagonal close packed.svg
Thermal expansion 6.2 µm/(m⋅K)
Thermal conductivity 48.0 W/(m⋅K)
Electrical resistivity 193 nΩ⋅m(at 20 °C)
Magnetic ordering paramagnetic [3]
Molar magnetic susceptibility +67.6×10−6 cm3/mol(293 K) [4]
Young's modulus 463 GPa
Shear modulus 178 GPa
Bulk modulus 370 GPa
Speed of sound thin rod4700 m/s(at 20 °C)
Poisson ratio 0.30
Mohs hardness 7.0
Vickers hardness 1350–7850 MPa
Brinell hardness 1320–2500 MPa
CAS Number 7440-15-5
History
Namingafter the river Rhine (German: Rhein)
Discovery Masataka Ogawa (1908)
First isolationMasataka Ogawa(1919)
Named by Walter Noddack, Ida Noddack, Otto Berg (1925)
Isotopes of rhenium
Main isotopes [5] Decay
abun­dance half-life (t1/2) mode pro­duct
185Re37.4% stable
186Resynth3.7185 d β 186Os
ε 186W
186mResynth2×105 y IT 186Re
β186Os
187Re62.6%4.12×1010 yβ 187Os
Symbol category class.svg  Category: Rhenium
| references

Rhenium is a chemical element; it has symbol Re and atomic number 75. It is a silvery-gray, heavy, third-row transition metal in group 7 of the periodic table. With an estimated average concentration of 1 part per billion (ppb), rhenium is one of the rarest elements in the Earth's crust. It has the third-highest melting point and second-highest boiling point of any element at 5869 K. [6] It resembles manganese and technetium chemically and is mainly obtained as a by-product of the extraction and refinement of molybdenum and copper ores. It shows in its compounds a wide variety of oxidation states ranging from −1 to +7.

Contents

Rhenium was originally discovered by Masataka Ogawa in 1908, but he mistakenly assigned it as element 43 rather than element 75 and named it nipponium. It was rediscovered by Walter Noddack, Ida Tacke and Otto Berg in 1925, [7] who gave it its present name. It was named after the river Rhine in Europe, from which the earliest samples had been obtained and worked commercially. [8]

Nickel-based superalloys of rhenium are used in combustion chambers, turbine blades, and exhaust nozzles of jet engines. These alloys contain up to 6% rhenium, making jet engine construction the largest single use for the element. The second-most important use is as a catalyst: it is an excellent catalyst for hydrogenation and isomerization, and is used for example in catalytic reforming of naphtha for use in gasoline (rheniforming process). Because of the low availability relative to demand, it is expensive, with price reaching an all-time high in 2008/2009 of US$10,600 per kilogram (US$4,800 per pound). Due to increases in recycling and a drop in demand for rhenium in catalysts, the price had dropped to US$2,844 per kilogram (US$1,290 per pound) as of July 2018. [9]

History

In 1908, Japanese chemist Masataka Ogawa announced that he had discovered the 43rd element and named it nipponium (Np) after Japan (Nippon in Japanese). In fact, he had found element 75 (rhenium) instead of element 43: both elements are in the same group of the periodic table. [10] [11] Ogawa's work was often incorrectly cited, because some of his key results were published only in Japanese; it is likely that his insistence on searching for element 43 prevented him from considering that he might have found element 75 instead. Just before Ogawa's death in 1930, Kenjiro Kimura analysed Ogawa's sample by X-ray spectroscopy at the Imperial University of Tokyo, and said to a friend that "it was beautiful rhenium indeed". He did not reveal this publicly, because under the Japanese university culture before World War II it was frowned upon to point out the mistakes of one's seniors, but the evidence became known to some Japanese news media regardless. As time passed with no repetitions of the experiments or new work on nipponium, Ogawa's claim faded away. [11] The symbol Np was later used for the element neptunium, and the name "nihonium", also named after Japan, along with symbol Nh, was later used for element 113. Element 113 was also discovered by a team of Japanese scientists and was named in respectful homage to Ogawa's work. [12] Today, Ogawa's claim is widely accepted as having been the discovery of element 75 in hindsight. [11]

Rhenium (Latin : Rhenus meaning: "Rhine") [13] received its current name when it was rediscovered by Walter Noddack, Ida Noddack, and Otto Berg in Germany. In 1925 they reported that they had detected the element in platinum ore and in the mineral columbite. They also found rhenium in gadolinite and molybdenite. [14] In 1928 they were able to extract 1 g of the element by processing 660 kg of molybdenite. [15] It was estimated in 1968 that 75% of the rhenium metal in the United States was used for research and the development of refractory metal alloys. It took several years from that point before the superalloys became widely used. [16] [17]

The original mischaracterization by Ogawa in 1908 and final work in 1925 makes rhenium perhaps the last stable element to be understood. Hafnium was discovered in 1923 [18] and all other new elements discovered since then, such as francium, are radioactive. [19]

Characteristics

Rhenium is a silvery-white metal with one of the highest melting points of all elements, exceeded by only tungsten. (At standard pressure carbon sublimes rather than melts, though its sublimation point is comparable to the melting points of tungsten and rhenium.) It also has one of the highest boiling points of all elements, and the highest among stable elements. It is also one of the densest, exceeded only by platinum, iridium and osmium. Rhenium has a hexagonal close-packed crystal structure, with lattice parameters a = 276.1 pm and c = 445.6 pm. [20]

Its usual commercial form is a powder, but this element can be consolidated by pressing and sintering in a vacuum or hydrogen atmosphere. This procedure yields a compact solid having a density above 90% of the density of the metal. When annealed this metal is very ductile and can be bent, coiled, or rolled. [21] Rhenium-molybdenum alloys are superconductive at 10 K; tungsten-rhenium alloys are also superconductive [22] around 4–8 K, depending on the alloy. Rhenium metal superconducts at 1.697±0.006 K. [23] [24]

In bulk form and at room temperature and atmospheric pressure, the element resists alkalis, sulfuric acid, hydrochloric acid, nitric acid, and aqua regia. It will however, react with nitric acid upon heating. [25]

Isotopes

Rhenium has one stable isotope, rhenium-185, which nevertheless occurs in minority abundance, a situation found only in two other elements (indium and tellurium). Naturally occurring rhenium is only 37.4% 185Re, and 62.6% 187Re, which is unstable but has a very long half-life (≈1010 years). A kilogram of natural rhenium emits 1.07 MBq of radiation due to the presence of this isotope. This lifetime can be greatly affected by the charge state of the rhenium atom. [26] [27] The beta decay of 187Re is used for rhenium–osmium dating of ores. The available energy for this beta decay (2.6  keV) is the second lowest known among all radionuclides, only behind the decay from 115In to excited 115Sn* (0.147 keV). [28] The isotope rhenium-186m is notable as being one of the longest lived metastable isotopes with a half-life of around 200,000 years. There are 33 other unstable isotopes that have been recognized, ranging from 160Re to 194Re, the longest-lived of which is 183Re with a half-life of 70 days. [29]

Compounds

Rhenium compounds are known for all the oxidation states between −3 and +7 except −2. The oxidation states +7, +4, and +3 are the most common. [30] Rhenium is most available commercially as salts of perrhenate, including sodium and ammonium perrhenates. These are white, water-soluble compounds. [31] Tetrathioperrhenate anion [ReS4] is possible. [32]

Halides and oxyhalides

The most common rhenium chlorides are ReCl6, ReCl5, ReCl4, and ReCl3. [33] The structures of these compounds often feature extensive Re-Re bonding, which is characteristic of this metal in oxidation states lower than VII. Salts of [Re2Cl8]2− feature a quadruple metal-metal bond. Although the highest rhenium chloride features Re(VI), fluorine gives the d0 Re(VII) derivative rhenium heptafluoride. Bromides and iodides of rhenium are also well known, including rhenium pentabromide and rhenium tetraiodide.

Like tungsten and molybdenum, with which it shares chemical similarities, rhenium forms a variety of oxyhalides. The oxychlorides are most common, and include ReOCl4, ReOCl3.

Oxides and sulfides

Perrhenic acid (H4Re2O9) adopts an unconventional structure. Perrhenic-acid-3D-balls.png
Perrhenic acid (H4Re2O9) adopts an unconventional structure.

The most common oxide is the volatile yellow Re2O7. The red rhenium trioxide ReO3 adopts a perovskite-like structure. Other oxides include Re2O5, ReO2, and Re2O3. [33] The sulfides are ReS2 and Re2S7. Perrhenate salts can be converted to tetrathioperrhenate by the action of ammonium hydrosulfide. [34]

Other compounds

Rhenium diboride (ReB2) is a hard compound having a hardness similar to that of tungsten carbide, silicon carbide, titanium diboride or zirconium diboride. [35]

Organorhenium compounds

Dirhenium decacarbonyl is the most common entry to organorhenium chemistry. Its reduction with sodium amalgam gives Na[Re(CO)5] with rhenium in the formal oxidation state −1. [36] Dirhenium decacarbonyl can be oxidised with bromine to bromopentacarbonylrhenium(I): [37]

Re2(CO)10 + Br2 → 2 Re(CO)5Br

Reduction of this pentacarbonyl with zinc and acetic acid gives pentacarbonylhydridorhenium: [38]

Re(CO)5Br + Zn + HOAc → Re(CO)5H + ZnBr(OAc)

Methylrhenium trioxide ("MTO"), CH3ReO3 is a volatile, colourless solid has been used as a catalyst in some laboratory experiments. It can be prepared by many routes, a typical method is the reaction of Re2O7 and tetramethyltin:

Re2O7 + (CH3)4Sn → CH3ReO3 + (CH3)3SnOReO3

Analogous alkyl and aryl derivatives are known. MTO catalyses for the oxidations with hydrogen peroxide. Terminal alkynes yield the corresponding acid or ester, internal alkynes yield diketones, and alkenes give epoxides. MTO also catalyses the conversion of aldehydes and diazoalkanes into an alkene. [39]

Nonahydridorhenate

Structure of ReH
9. Nonahydridorhenate-3D-balls.png
Structure of ReH
9
.

A distinctive derivative of rhenium is nonahydridorhenate, originally thought to be the rhenide anion, Re, but actually containing the ReH2−
9
anion in which the oxidation state of rhenium is +7.

Occurrence

Molybdenite Molybdenit 1.jpg
Molybdenite

Rhenium is one of the rarest elements in Earth's crust with an average concentration of 1 ppb; [33] other sources quote the number of 0.5 ppb making it the 77th most abundant element in Earth's crust. [40] Rhenium is probably not found free in nature (its possible natural occurrence is uncertain), but occurs in amounts up to 0.2% [33] in the mineral molybdenite (which is primarily molybdenum disulfide), the major commercial source, although single molybdenite samples with up to 1.88% have been found. [41] Chile has the world's largest rhenium reserves, part of the copper ore deposits, and was the leading producer as of 2005. [42] It was only recently that the first rhenium mineral was found and described (in 1994), a rhenium sulfide mineral (ReS2) condensing from a fumarole on Kudriavy volcano, Iturup island, in the Kuril Islands. [43] Kudriavy discharges up to 20–60 kg rhenium per year mostly in the form of rhenium disulfide. [44] [45] Named rheniite, this rare mineral commands high prices among collectors. [46]

Production

Ammonium perrhenate Ammonium perrhenate.jpg
Ammonium perrhenate

Approximately 80% of rhenium is extracted from porphyry molybdenum deposits. [47] Some ores contain 0.001% to 0.2% rhenium. [33] Roasting the ore volatilizes rhenium oxides. [41] Rhenium(VII) oxide and perrhenic acid readily dissolve in water; they are leached from flue dusts and gasses and extracted by precipitating with potassium or ammonium chloride as the perrhenate salts, and purified by recrystallization. [33] Total world production is between 40 and 50 tons/year; the main producers are in Chile, the United States, Peru, and Poland. [48] Recycling of used Pt-Re catalyst and special alloys allow the recovery of another 10 tons per year. Prices for the metal rose rapidly in early 2008, from $1000–$2000 per kg in 2003–2006 to over $10,000 in February 2008. [49] [50] The metal form is prepared by reducing ammonium perrhenate with hydrogen at high temperatures: [31]

2 NH4ReO4 + 7 H2 → 2 Re + 8 H2O + 2 NH3

There are technologies for the associated extraction of rhenium from productive solutions of underground leaching of uranium ores. [51]

Applications

The Pratt & Whitney F-100 engine uses rhenium-containing second-generation superalloys Engine.f15.arp.750pix.jpg
The Pratt & Whitney F-100 engine uses rhenium-containing second-generation superalloys

Rhenium is added to high-temperature superalloys that are used to make jet engine parts, using 70% of the worldwide rhenium production. [52] Another major application is in platinum–rhenium catalysts, which are primarily used in making lead-free, high-octane gasoline. [53]

Alloys

The nickel-based superalloys have improved creep strength with the addition of rhenium. The alloys normally contain 3% or 6% of rhenium. [54] Second-generation alloys contain 3%; these alloys were used in the engines for the F-15 and F-16, whereas the newer single-crystal third-generation alloys contain 6% of rhenium; they are used in the F-22 and F-35 engines. [53] [55] Rhenium is also used in the superalloys, such as CMSX-4 (2nd gen) and CMSX-10 (3rd gen) that are used in industrial gas turbine engines like the GE 7FA. Rhenium can cause superalloys to become microstructurally unstable, forming undesirable topologically close packed (TCP) phases. In 4th- and 5th-generation superalloys, ruthenium is used to avoid this effect. Among others the new superalloys are EPM-102 (with 3% Ru) and TMS-162 (with 6% Ru), [56] as well as TMS-138 [57] and TMS-174. [58] [59]

CFM International CFM56 jet engine with blades made with 3% rhenium CFM56 P1220759.jpg
CFM International CFM56 jet engine with blades made with 3% rhenium

For 2006, the consumption is given as 28% for General Electric, 28% Rolls-Royce plc and 12% Pratt & Whitney, all for superalloys, whereas the use for catalysts only accounts for 14% and the remaining applications use 18%. [52] In 2006, 77% of rhenium consumption in the United States was in alloys. [53] The rising demand for military jet engines and the constant supply made it necessary to develop superalloys with a lower rhenium content. For example, the newer CFM International CFM56 high-pressure turbine (HPT) blades will use Rene N515 with a rhenium content of 1.5% instead of Rene N5 with 3%. [60] [61]

Rhenium improves the properties of tungsten. Tungsten-rhenium alloys are more ductile at low temperature, allowing them to be more easily machined. The high-temperature stability is also improved. The effect increases with the rhenium concentration, and therefore tungsten alloys are produced with up to 27% of Re, which is the solubility limit. [62] Tungsten-rhenium wire was originally created in efforts to develop a wire that was more ductile after recrystallization. This allows the wire to meet specific performance objectives, including superior vibration resistance, improved ductility, and higher resistivity. [63] One application for the tungsten-rhenium alloys is X-ray sources. The high melting point of both elements, together with their high atomic mass, makes them stable against the prolonged electron impact. [64] Rhenium tungsten alloys are also applied as thermocouples to measure temperatures up to 2200 °C. [65]

The high temperature stability, low vapor pressure, good wear resistance and ability to withstand arc corrosion of rhenium are useful in self-cleaning electrical contacts. In particular, the discharge that occurs during electrical switching oxidizes the contacts. However, rhenium oxide Re2O7 is volatile (sublimes at ~360 °C) and therefore is removed during the discharge. [52]

Rhenium has a high melting point and a low vapor pressure similar to tantalum and tungsten. Therefore, rhenium filaments exhibit a higher stability if the filament is operated not in vacuum, but in oxygen-containing atmosphere. [66] Those filaments are widely used in mass spectrometers, ion gauges [67] and photoflash lamps in photography. [68]

Catalysts

Rhenium in the form of rhenium-platinum alloy is used as catalyst for catalytic reforming, which is a chemical process to convert petroleum refinery naphthas with low octane ratings into high-octane liquid products. Worldwide, 30% of catalysts used for this process contain rhenium. [69] The olefin metathesis is the other reaction for which rhenium is used as catalyst. Normally Re2O7 on alumina is used for this process. [70] Rhenium catalysts are very resistant to chemical poisoning from nitrogen, sulfur and phosphorus, and so are used in certain kinds of hydrogenation reactions. [21] [71] [72]

Other uses

The isotopes 186Re and 188Re are radioactive and are used for treatment of liver cancer. They both have similar penetration depth in tissue (5 mm for 186Re and 11 mm for 188Re), but 186Re has the advantage of a longer half life (90 hours vs. 17 hours). [73] [74]

188Re is also being used experimentally in a novel treatment of pancreatic cancer where it is delivered by means of the bacterium Listeria monocytogenes. [75] The 188Re isotope is also used for the rhenium-SCT (skin cancer therapy). The treatment uses the isotope's properties as a beta emitter for brachytherapy in the treatment of basal cell carcinoma and squamous cell carcinoma of the skin. [76]

Related by periodic trends, rhenium has a similar chemistry to that of technetium; work done to label rhenium onto target compounds can often be translated to technetium. This is useful for radiopharmacy, where it is difficult to work with technetium – especially the technetium-99m isotope used in medicine – due to its expense and short half-life. [73] [77]

Precautions

Very little is known about the toxicity of rhenium and its compounds because they are used in very small amounts. Soluble salts, such as the rhenium halides or perrhenates, could be hazardous due to elements other than rhenium or due to rhenium itself. [78] Only a few compounds of rhenium have been tested for their acute toxicity; two examples are potassium perrhenate and rhenium trichloride, which were injected as a solution into rats. The perrhenate had an LD50 value of 2800 mg/kg after seven days (this is very low toxicity, similar to that of table salt) and the rhenium trichloride showed LD50 of 280 mg/kg. [79]

Related Research Articles

<span class="mw-page-title-main">Hafnium</span> Chemical element, symbol Hf and atomic number 72

Hafnium is a chemical element; it has symbol Hf and atomic number 72. A lustrous, silvery gray, tetravalent transition metal, hafnium chemically resembles zirconium and is found in many zirconium minerals. Its existence was predicted by Dmitri Mendeleev in 1869, though it was not identified until 1922, by Dirk Coster and George de Hevesy. making it one of the last two stable elements to be discovered. Hafnium is named after Hafnia, the Latin name for Copenhagen, where it was discovered.

<span class="mw-page-title-main">Osmium</span> Chemical element, symbol Os and atomic number 76

Osmium is a chemical element; it has symbol Os and atomic number 76. It is a hard, brittle, bluish-white transition metal in the platinum group that is found as a trace element in alloys, mostly in platinum ores. Osmium is the densest naturally occurring element. When experimentally measured using X-ray crystallography, it has a density of 22.59 g/cm3. Manufacturers use its alloys with platinum, iridium, and other platinum-group metals to make fountain pen nib tipping, electrical contacts, and in other applications that require extreme durability and hardness.

<span class="mw-page-title-main">Ruthenium</span> Chemical element, symbol Ru and atomic number 44

Ruthenium is a chemical element; it has symbol Ru and atomic number 44. It is a rare transition metal belonging to the platinum group of the periodic table. Like the other metals of the platinum group, ruthenium is inert to most other chemicals. Karl Ernst Claus, a Russian-born scientist of Baltic-German ancestry, discovered the element in 1844 at Kazan State University and named ruthenium in honor of Russia. Ruthenium is usually found as a minor component of platinum ores; the annual production has risen from about 19 tonnes in 2009 to some 35.5 tonnes in 2017. Most ruthenium produced is used in wear-resistant electrical contacts and thick-film resistors. A minor application for ruthenium is in platinum alloys and as a chemistry catalyst. A new application of ruthenium is as the capping layer for extreme ultraviolet photomasks. Ruthenium is generally found in ores with the other platinum group metals in the Ural Mountains and in North and South America. Small but commercially important quantities are also found in pentlandite extracted from Sudbury, Ontario, and in pyroxenite deposits in South Africa.

<span class="mw-page-title-main">Technetium</span> Chemical element, symbol Tc and atomic number 43

Technetium is a chemical element; it has symbol Tc and atomic number 43. It is the lightest element whose isotopes are all radioactive. Technetium and promethium are the only radioactive elements whose neighbours in the sense of atomic number are both stable. All available technetium is produced as a synthetic element. Naturally occurring technetium is a spontaneous fission product in uranium ore and thorium ore, or the product of neutron capture in molybdenum ores. This silvery gray, crystalline transition metal lies between manganese and rhenium in group 7 of the periodic table, and its chemical properties are intermediate between those of both adjacent elements. The most common naturally occurring isotope is 99Tc, in traces only.

<span class="mw-page-title-main">Tungsten</span> Chemical element, symbol W and atomic number 74

Tungsten is a chemical element; it has symbol W and atomic number 74. Tungsten is a rare metal found naturally on Earth almost exclusively as compounds with other elements. It was identified as a new element in 1781 and first isolated as a metal in 1783. Its important ores include scheelite and wolframite, the latter lending the element its alternative name.

A period 5 element is one of the chemical elements in the fifth row of the periodic table of the chemical elements. The periodic table is laid out in rows to illustrate recurring (periodic) trends in the chemical behaviour of the elements as their atomic number increases: a new row is begun when chemical behaviour begins to repeat, meaning that elements with similar behaviour fall into the same vertical columns. The fifth period contains 18 elements, beginning with rubidium and ending with xenon. As a rule, period 5 elements fill their 5s shells first, then their 4d, and 5p shells, in that order; however, there are exceptions, such as rhodium.

A period 6 element is one of the chemical elements in the sixth row (or period) of the periodic table of the chemical elements, including the lanthanides. The periodic table is laid out in rows to illustrate recurring (periodic) trends in the chemical behaviour of the elements as their atomic number increases: a new row is begun when chemical behaviour begins to repeat, meaning that elements with similar behaviour fall into the same vertical columns. The sixth period contains 32 elements, tied for the most with period 7, beginning with caesium and ending with radon. Lead is currently the last stable element; all subsequent elements are radioactive. For bismuth, however, its only primordial isotope, 209Bi, has a half-life of more than 1019 years, over a billion times longer than the current age of the universe. As a rule, period 6 elements fill their 6s shells first, then their 4f, 5d, and 6p shells, in that order; however, there are exceptions, such as gold.

<span class="mw-page-title-main">Group 7 element</span> Group of chemical elements

Group 7, numbered by IUPAC nomenclature, is a group of elements in the periodic table. It contains manganese (Mn), technetium (Tc), rhenium (Re) and bohrium (Bh). This group lies in the d-block of the periodic table, and are hence transition metals. This group is sometimes called the manganese group or manganese family after its lightest member; however, the group itself has not acquired a trivial name because it belongs to the broader grouping of the transition metals.

Refractory metals are a class of metals that are extraordinarily resistant to heat and wear. The expression is mostly used in the context of materials science, metallurgy and engineering. The definition of which elements belong to this group differs. The most common definition includes five elements: two of the fifth period and three of the sixth period. They all share some properties, including a melting point above 2000 °C and high hardness at room temperature. They are chemically inert and have a relatively high density. Their high melting points make powder metallurgy the method of choice for fabricating components from these metals. Some of their applications include tools to work metals at high temperatures, wire filaments, casting molds, and chemical reaction vessels in corrosive environments. Partly due to the high melting point, refractory metals are stable against creep deformation to very high temperatures.

Molybdenum trioxide describes a family of inorganic compounds with the formula MoO3(H2O)n where n = 0, 1, 2. The anhydrous compound is produced on the largest scale of any molybdenum compound since it is the main intermediate produced when molybdenum ores are purified. The anhydrous oxide is a precursor to molybdenum metal, an important alloying agent. It is also an important industrial catalyst. It is a yellow solid, although impure samples can appear blue or green.

<span class="mw-page-title-main">Perrhenic acid</span> Chemical compound

Perrhenic acid is the chemical compound with the formula Re2O7(H2O)2. It is obtained by evaporating aqueous solutions of Re2O7. Conventionally, perrhenic acid is considered to have the formula HReO4, and a species of this formula forms when rhenium(VII) oxide sublimes in the presence of water or steam. When a solution of Re2O7 is kept for a period of months, it breaks down and crystals of HReO4·H2O are formed, which contain tetrahedral ReO−4. For most purposes, perrhenic acid and rhenium(VII) oxide are used interchangeably. Rhenium can be dissolved in nitric or concentrated sulfuric acid to produce perrhenic acid.

<span class="mw-page-title-main">Ida Noddack</span> German chemist (1896–1978)

Ida Noddack, néeTacke, was a German chemist and physicist. In 1934 she was the first to mention the idea later named nuclear fission. With her husband Walter Noddack, and Otto Berg, she discovered element 75, rhenium. She was nominated three times for the Nobel Prize in Chemistry.

<span class="mw-page-title-main">Ammonium perrhenate</span> Chemical compound

Ammonium perrhenate (APR) is the ammonium salt of perrhenic acid, NH4ReO4. It is the most common form in which rhenium is traded. It is a white salt; soluble in ethanol and water, and mildly soluble in NH4Cl. It was first described soon after the discovery of rhenium.

<span class="mw-page-title-main">Rhenium(VII) oxide</span> Chemical compound

Rhenium(VII) oxide is the inorganic compound with the formula Re2O7. This yellowish solid is the anhydride of HOReO3. Perrhenic acid, Re2O7·2H2O, is closely related to Re2O7. Re2O7 is the raw material for all rhenium compounds, being the volatile fraction obtained upon roasting the host ore.

<span class="mw-page-title-main">Dirhenium decacarbonyl</span> Chemical compound

Dirhenium decacarbonyl is the inorganic compound with the chemical formula Re2(CO)10. Commercially available, it is used as a starting point for the synthesis of many rhenium carbonyl complexes. It was first reported in 1941 by Walter Hieber, who prepared it by reductive carbonylation of rhenium. The compound consists of a pair of square pyramidal Re(CO)5 units joined via a Re-Re bond, which produces a homoleptic carbonyl complex.

<span class="mw-page-title-main">Masataka Ogawa</span> Japanese chemist

Masataka Ogawa was a Japanese chemist mainly known for the claimed discovery of element 43, which he named nipponium. In fact, he had discovered, but misidentified, element 75.

The perrhenate ion is the anion with the formula ReO
4
, or a compound containing this ion. The perrhenate anion is tetrahedral, being similar in size and shape to perchlorate and the valence isoelectronic permanganate. The perrhenate anion is stable over a broad pH range and can be precipitated from solutions with the use of organic cations. At normal pH, perrhenate exists as metaperrhenate, but at high pH mesoperrhenate forms. Perrhenate, like its conjugate acid perrhenic acid, features rhenium in the oxidation state of +7 with a d0 configuration. Solid perrhenate salts takes on the color of the cation.

<span class="mw-page-title-main">Otto Berg (scientist)</span> German scientist (1873–1939)

Otto Berg was a German scientist. He is one of the scientists credited with discovering rhenium, the last element to be discovered having a stable isotope.

Organorhenium chemistry describes the compounds with Re−C bonds. Because rhenium is a rare element, relatively few applications exist, but the area has been a rich source of concepts and a few useful catalysts.

Rhenium compounds are compounds formed by the transition metal rhenium (Re). Rhenium can form in many oxidation states, and compounds are known for every oxidation state from -3 to +7 except -2, although the oxidation states +7, +4, and +3 are the most common. Rhenium is most available commercially as salts of perrhenate, including sodium and ammonium perrhenates. These are white, water-soluble compounds. The tetrathioperrhenate anion [ReS4] is possible.

References

  1. "Standard Atomic Weights: Rhenium". CIAAW. 1973.
  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. 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.
  4. Weast, Robert (1984). CRC, Handbook of Chemistry and Physics. Boca Raton, Florida: Chemical Rubber Company Publishing. pp. E110. ISBN   0-8493-0464-4.
  5. Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S.; Audi, G. (2021). "The NUBASE2020 evaluation of nuclear properties" (PDF). Chinese Physics C. 45 (3): 030001. doi:10.1088/1674-1137/abddae.
  6. Zhang, Yiming (2011-01-11). "Corrected Values for Boiling Points and Enthalpies of Vaporization of Elements in Handbooks". Journal of Chemical & Engineering Data. 56.
  7. "Die Ekamangane". Naturwissenschaften (in German). 13 (26): 567–574. 1925-06-01. Bibcode:1925NW.....13..567.. doi:10.1007/BF01558746. ISSN   1432-1904. S2CID   32974087.
  8. "From Hydrogen to Darmstadtium & More". American Chemical Society. 2003. p. 144.
  9. "BASF Catalysts - Metal Prices". apps.catalysts.basf.com.
  10. Yoshihara, H. K. (2004). "Discovery of a new element 'nipponiumʼ: re-evaluation of pioneering works of Masataka Ogawa and his son Eijiro Ogawa". Spectrochimica Acta Part B: Atomic Spectroscopy. 59 (8): 1305–1310. Bibcode:2004AcSpe..59.1305Y. doi:10.1016/j.sab.2003.12.027.
  11. 1 2 3 Hisamatsu, Yoji; Egashira, Kazuhiro; Maeno, Yoshiteru (2022). "Ogawa's nipponium and its re-assignment to rhenium". Foundations of Chemistry. 24: 15–57. doi: 10.1007/s10698-021-09410-x .
  12. Öhrström, Lars; Reedijk, Jan (28 November 2016). "Names and symbols of the elements with atomic numbers 113, 115, 117 and 118 (IUPAC Recommendations 2016)" (PDF). Pure Appl. Chem. 88 (12): 1225–1229. doi:10.1515/pac-2016-0501. hdl: 1887/47427 . S2CID   99429711 . Retrieved 22 April 2017.
  13. Tilgner, Hans Georg (2000). Forschen Suche und Sucht (in German). Books on Demand. ISBN   978-3-89811-272-7.
  14. Noddack, W.; Tacke, I.; Berg, O. (1925). "Die Ekamangane". Naturwissenschaften. 13 (26): 567–574. Bibcode:1925NW.....13..567.. doi:10.1007/BF01558746. S2CID   32974087.
  15. Noddack, W.; Noddack, I. (1929). "Die Herstellung von einem Gram Rhenium". Zeitschrift für Anorganische und Allgemeine Chemie (in German). 183 (1): 353–375. doi:10.1002/zaac.19291830126.
  16. Committee On Technical Aspects Of Critical And Strategic Material, National Research Council (U.S.) (1968). Trends in usage of rhenium: Report. pp. 4–5.
  17. Savitskiĭ, Evgeniĭ Mikhaĭlovich; Tulkina, Mariia Aronovna; Povarova, Kira Borisovna (1970). Rhenium alloys.
  18. "Two Danes Discover New Element, Hafnium Detect It by Means of Spectrum Analysis of Ore Containing Zirconium", The New York Times, January 20, 1923, p. 4
  19. "Rhenium: Statistics and Information". Minerals Information. United States Geological Survey. 2011. Retrieved 2011-05-25.
  20. Liu, L. G.; Takahashi, T.; Bassett, W. A. (1970). "Effect of pressure and temperature on lattice parameters of rhenium". Journal of Physics and Chemistry of Solids. 31 (6): 1345–1351. Bibcode:1970JPCS...31.1345L. doi:10.1016/0022-3697(70)90138-1.
  21. 1 2 Hammond, C. R. (2004). "The Elements" . Handbook of Chemistry and Physics (81st ed.). CRC press. ISBN   978-0-8493-0485-9.
  22. Neshpor, V. S.; Novikov, V. I.; Noskin, V. A.; Shalyt, S. S. (1968). "Superconductivity of Some Alloys of the Tungsten-rhenium-carbon System". Soviet Physics JETP. 27: 13. Bibcode:1968JETP...27...13N.
  23. Haynes, William M., ed. (2011). CRC Handbook of Chemistry and Physics (92nd ed.). CRC Press. p. 12.60. ISBN   978-1439855119.
  24. Daunt, J. G.; Lerner, E. "The Properties of Superconducting Mo-Re Alloys". Defense Technical Information Center. Archived from the original on 2017-02-06.
  25. "Rhenium - A METAL WITHOUT WHICH THERE WOULdn't BE GASOLINE!". YouTube .
  26. Johnson, Bill (1993). "How to Change Nuclear Decay Rates". math.ucr.edu. Retrieved 2009-02-21.
  27. Bosch, F.; Faestermann, T.; Friese, J.; et al. (1996). "Observation of bound-state β decay of fully ionized 187Re: 187Re-187Os Cosmochronometry". Physical Review Letters . 77 (26): 5190–5193. Bibcode:1996PhRvL..77.5190B. doi:10.1103/PhysRevLett.77.5190. PMID   10062738.
  28. Belli, P.; Bernabei, R.; Danevich, F. A.; Incicchitti, A.; Tretyak, V. I. (2019). "Experimental searches for rare alpha and beta decays". The European Physical Journal A. 55 (8). Springer Science and Business Media LLC: 140. arXiv: 1908.11458 . Bibcode:2019EPJA...55..140B. doi:10.1140/epja/i2019-12823-2. ISSN   1434-6001.
  29. Audi, G.; Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S. (2017). "The NUBASE2016 evaluation of nuclear properties" (PDF). Chinese Physics C. 41 (3): 030001. Bibcode:2017ChPhC..41c0001A. doi:10.1088/1674-1137/41/3/030001.
  30. Housecroft, Catherine E.; Sharpe, Alan G. (2018). Inorganic Chemistry (5th ed.). Pearson Prentice-Hal. p. 829. ISBN   978-1292-13414-7.
  31. 1 2 Glemser, O. (1963) "Ammonium Perrhenate" in Handbook of Preparative Inorganic Chemistry, 2nd ed., G. Brauer (ed.), Academic Press, NY., Vol. 1, pp. 1476–85.
  32. Goodman, JT; Rauchfuss, TB (2002). "Useful Reagents and Ligands". Inorganic Syntheses. Inorganic Syntheses. Vol. 33. pp. 107–110. doi:10.1002/0471224502.ch2. ISBN   0471208256.
  33. 1 2 3 4 5 6 Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. ISBN   978-0-08-037941-8.
  34. Goodman, J. T.; Rauchfuss, T. B. (2002). "Useful Reagents and Ligands". Inorganic Syntheses. Vol. 33. pp. 107–110. doi:10.1002/0471224502.ch2. ISBN   9780471208259.
  35. Qin, Jiaqian; He, Duanwei; Wang, Jianghua; Fang, Leiming; Lei, Li; Li, Yongjun; Hu, Juan; Kou, Zili; Bi, Yan (2008). "Is Rhenium Diboride a Superhard Material?". Advanced Materials. 20 (24): 4780–4783. Bibcode:2008AdM....20.4780Q. doi:10.1002/adma.200801471. S2CID   98327405.
  36. Breimair, Josef; Steimann, Manfred; Wagner, Barbara; Beck, Wolfgang (1990). "Nucleophile Addition von Carbonylmetallaten an kationische Alkin-Komplexe [CpL2M(η2-RC≡CR)]+ (M = Ru, Fe): μ-η1:η1-Alkin-verbrückte Komplexe". Chemische Berichte. 123: 7. doi:10.1002/cber.19901230103.
  37. Schmidt, Steven P.; Trogler, William C.; Basolo, Fred (1990). "Pentacarbonylrhenium Halides". Inorganic Syntheses. Vol. 28. pp. 154–159. doi:10.1002/9780470132593.ch42. ISBN   978-0-470-13259-3.
  38. Michael A. Urbancic; John R. Shapley (1990). "Pentacarbonylhydridorhenium". Inorganic Syntheses. Vol. 28. pp. 165–168. doi:10.1002/9780470132593.ch43. ISBN   978-0-470-13259-3.
  39. Hudson, A. (2002) “Methyltrioxorhenium” in Encyclopedia of Reagents for Organic Synthesis. John Wiley & Sons: New York, ISBN   9780470842898, doi : 10.1002/047084289X.
  40. Emsley, John (2001). "Rhenium". Nature's Building Blocks: An A-Z Guide to the Elements. Oxford, England, UK: Oxford University Press. pp.  358–360. ISBN   978-0-19-850340-8.
  41. 1 2 Rouschias, George (1974). "Recent advances in the chemistry of rhenium". Chemical Reviews. 74 (5): 531. doi:10.1021/cr60291a002.
  42. Anderson, Steve T. "2005 Minerals Yearbook: Chile" (PDF). United States Geological Survey . Retrieved 2008-10-26.
  43. Korzhinsky, M. A.; Tkachenko, S. I.; Shmulovich, K. I.; Taran Y. A.; Steinberg, G. S. (2004-05-05). "Discovery of a pure rhenium mineral at Kudriavy volcano". Nature . 369 (6475): 51–52. Bibcode:1994Natur.369...51K. doi:10.1038/369051a0. S2CID   4344624.
  44. Kremenetsky, A. A.; Chaplygin, I. V. (2010). "Concentration of rhenium and other rare metals in gases of the Kudryavy Volcano (Iturup Island, Kurile Islands)". Doklady Earth Sciences. 430 (1): 114. Bibcode:2010DokES.430..114K. doi:10.1134/S1028334X10010253. S2CID   140632604.
  45. Tessalina, S.; Yudovskaya, M.; Chaplygin, I.; Birck, J.; Capmas, F. (2008). "Sources of unique rhenium enrichment in fumaroles and sulphides at Kudryavy volcano". Geochimica et Cosmochimica Acta. 72 (3): 889. Bibcode:2008GeCoA..72..889T. doi:10.1016/j.gca.2007.11.015.
  46. "The Mineral Rheniite". Amethyst Galleries.
  47. John, D. A.; Taylor, R. D. (2016). "Chapter 7: By-Products of Porphyry Copper and Molybdenum Deposits". In Philip L. Verplanck and Murray W. Hitzman (ed.). Rare earth and critical elements in ore deposits. Vol. 18. pp. 137–164. doi:10.5382/Rev.18.07.
  48. Magyar, Michael J. (January 2012). "Rhenium" (PDF). Mineral Commodity Summaries. U.S. Geological Survey. Retrieved 2013-09-04.
  49. "MinorMetal prices". minormetals.com. Retrieved 2008-02-17.
  50. Harvey, Jan (2008-07-10). "Analysis: Super hot metal rhenium may reach "platinum prices"". Reuters India. Retrieved 2008-10-26.
  51. Rudenko, A.A.; Troshkina, I.D.; Danileyko, V.V.; Barabanov, O.S.; Vatsura, F.Y. (2021). "Prospects for selective-and-advanced recovery of rhenium from pregnant solutions of in-situ leaching of uranium ores at Dobrovolnoye deposit". Gornye Nauki I Tekhnologii = Mining Science and Technology (Russia). 6 (3): 158–169. doi:10.17073/2500-0632-2021-3-158-169. S2CID   241476783.
  52. 1 2 3 Naumov, A. V. (2007). "Rhythms of rhenium". Russian Journal of Non-Ferrous Metals. 48 (6): 418–423. doi:10.3103/S1067821207060089. S2CID   137550564.
  53. 1 2 3 Magyar, Michael J. (April 2011). "2009 Mineral Yearbook: Rhenium" (PDF). United States Geological Survey.
  54. Bhadeshia, H. K. D. H. "Nickel Based Superalloys". University of Cambridge. Archived from the original on 2006-08-25. Retrieved 2008-10-17.
  55. Cantor, B.; Grant, Patrick Assender Hazel (2001). Aerospace Materials: An Oxford-Kobe Materials Text. CRC Press. pp. 82–83. ISBN   978-0-7503-0742-0.
  56. Bondarenko, Yu. A.; Kablov, E. N.; Surova, V. A.; Echin, A. B. (2006). "Effect of high-gradient directed crystallization on the structure and properties of rhenium-bearing single-crystal alloy". Metal Science and Heat Treatment. 48 (7–8): 360. Bibcode:2006MSHT...48..360B. doi:10.1007/s11041-006-0099-6. S2CID   136907279.
  57. "Fourth generation nickel base single crystal superalloy" (PDF).
  58. Koizumi, Yutaka; et al. "Development of a Next-Generation Ni-base Single Crystal Superalloy" (PDF). Proceedings of the International Gas Turbine Congress, Tokyo November 2–7, 2003.
  59. Walston, S.; Cetel, A.; MacKay, R.; O'Hara, K.; Duhl, D.; Dreshfield, R. "Joint Development of a Fourth Generation Single Crystal Superalloy" (PDF). Archived from the original (PDF) on 2006-10-15.
  60. Fink, Paul J.; Miller, Joshua L.; Konitzer, Douglas G. (2010). "Rhenium reduction—alloy design using an economically strategic element". JOM. 62 (1): 55. Bibcode:2010JOM....62a..55F. doi:10.1007/s11837-010-0012-z. S2CID   137007996.
  61. Konitzer, Douglas G. (September 2010). "Design in an Era of Constrained Resources". Archived from the original on 2011-07-25. Retrieved 2010-10-12.
  62. Lassner, Erik; Schubert, Wolf-Dieter (1999). Tungsten: properties, chemistry, technology of the element, alloys, and chemical compounds. Springer. p. 256. ISBN   978-0-306-45053-2.
  63. "Tungsten-Rhenium - Union City Filament". Union City Filament. Retrieved 2017-04-05.
  64. Cherry, Pam; Duxbury, Angela (1998). Practical radiotherapy physics and equipment. Cambridge University Press. p. 55. ISBN   978-1-900151-06-1.
  65. Asamoto, R.; Novak, P. E. (1968). "Tungsten-Rhenium Thermocouples for Use at High Temperatures". Review of Scientific Instruments. 39 (8): 1233. Bibcode:1968RScI...39.1233A. doi:10.1063/1.1683642.
  66. Blackburn, Paul E. (1966). "The Vapor Pressure of Rhenium". The Journal of Physical Chemistry. 70: 311–312. doi:10.1021/j100873a513.
  67. Earle, G. D.; Medikonduri, R.; Rajagopal, N.; Narayanan, V.; Roddy, P. A. (2005). "Tungsten-Rhenium Filament Lifetime Variability in Low Pressure Oxygen Environments". IEEE Transactions on Plasma Science. 33 (5): 1736–1737. Bibcode:2005ITPS...33.1736E. doi:10.1109/TPS.2005.856413. S2CID   26162679.
  68. Ede, Andrew (2006). The chemical element: a historical perspective. Greenwood Publishing Group. ISBN   978-0-313-33304-0.
  69. Ryashentseva, Margarita A. (1998). "Rhenium-containing catalysts in reactions of organic compounds". Russian Chemical Reviews. 67 (2): 157–177. Bibcode:1998RuCRv..67..157R. doi:10.1070/RC1998v067n02ABEH000390. S2CID   250866233.
  70. Mol, Johannes C. (1999). "Olefin metathesis over supported rhenium oxide catalysts". Catalysis Today. 51 (2): 289–299. doi:10.1016/S0920-5861(99)00051-6.
  71. Angelidis, T. N.; Rosopoulou, D. Tzitzios V. (1999). "Selective Rhenium Recovery from Spent Reforming Catalysts". Ind. Eng. Chem. Res. 38 (5): 1830–1836. doi:10.1021/ie9806242.
  72. Burch, Robert (1978). "The Oxidation State of Rhenium and Its Role in Platinum-Rhenium" (PDF). Platinum Metals Review. 22 (2): 57–60. doi:10.1595/003214078X2225760.
  73. 1 2 Dilworth, Jonathan R.; Parrott, Suzanne J. (1998). "The biomedical chemistry of technetium and rhenium". Chemical Society Reviews. 27: 43–55. doi:10.1039/a827043z.
  74. "The Tungsten-188 and Rhenium-188 Generator Information". Oak Ridge National Laboratory. 2005. Archived from the original on 2008-01-09. Retrieved 2008-02-03.
  75. Baker, Monya (22 April 2013). "Radioactive bacteria attack cancer". Nature. doi: 10.1038/nature.2013.12841 .
  76. Cipriani, Cesidio; Desantis, Maria; Dahlhoff, Gerhard; Brown, Shannon D.; Wendler, Thomas; Olmeda, Mar; Pietsch, Gunilla; Eberlein, Bernadette (2020-07-22). "Personalized irradiation therapy for NMSC by rhenium-188 skin cancer therapy: a long-term retrospective study". Journal of Dermatological Treatment. 33 (2): 969–975. doi: 10.1080/09546634.2020.1793890 . ISSN   0954-6634. PMID   32648530.
  77. Colton, R.; Peacock R. D. (1962). "An outline of technetium chemistry". Quarterly Reviews, Chemical Society. 16 (4): 299–315. doi:10.1039/QR9621600299.
  78. Emsley, J. (2003). "Rhenium". Nature's Building Blocks: An A-Z Guide to the Elements. Oxford, England, UK: Oxford University Press. pp.  358–361. ISBN   978-0-19-850340-8.
  79. Haley, Thomas J.; Cartwright, Frank D. (1968). "Pharmacology and toxicology of potassium perrhenate and rhenium trichloride". Journal of Pharmaceutical Sciences. 57 (2): 321–323. doi:10.1002/jps.2600570218. PMID   5641681.

Further reading