User:Marshallsumter/Radiation astronomy/Alloys

From Wikiversity
Jump to navigation Jump to search
File:Swarm-satellite.jpg
The "Swarm" satellites have been flying around Earth since Fall of 2013. Credit: Christoph Seidler, ESA/DTU.{{fairuse}}

"Three [Swarm] satellites of the European Space Agency (ESA) have measured the magnetic field of Earth more precisely than ever before."[1]

Alloys included here may be "a combination of two or more elements, at least one of which is a metal", an "admixture", anything "made by combining several things", a "substance made from any combination of ingredients" or "a substance made from the chemical combination of elements".

Explorations (Earth)

[edit | edit source]
Phosphate mining in Togo is a last phase of exploration geology. Credit: Alexandra Pugachevsky.{{free media}}

"Exploration geology is the single most important and very first phase of mining. It begins by identifying what mineral/minerals is/are to be exploited, their geological setting, approximate size of orebody required and potential areas. Once these factors are considered, funds are required to finance the exploration project. Usually exploration companies list on stock exchanges to raise the required capital. Exploration begins by firstly gathering any possible data available on the resource, area, local geology usually from the geological survey, from satellite imagery as well as previous scientific work. The next phase usually involves geotechnical prospecting which makes use of either seismic, electrical, magnetic, radioactive or density techniques. Once a suitable area has been found, holes are drilled and the core retrieved is logged and correlated against other logs to form a model of the orebody. Once sufficient holes have been drilled and the ore tested for qualities, feasibility studies and due diligence work can commence."[2]

Alloys

[edit | edit source]

Alloys are defined by a metallic bonding character.[3]

Def. a "metal that is a combination of two or more elements, at least one of which is a metal"[4] or an "admixture"[5], "instance of admixing [mingling], a mixing-in of something"[6] is called an alloy.

Def. anything "made by combining several things",[7] a "substance formed by chemical union [bonding][8] of two or more ingredients [elements][8] in definite proportions by weight",[9] a "substance made from any combination of ingredients",[8] "a substance made from the chemical combination of elements"[10] is called a compound.

Sulfur combines readily with iron to form iron sulfide, which is very brittle, creating weak spots in the steel.[11]

The physical properties, such as density, reactivity, Young's modulus of an alloy may not differ greatly from those of its base element, but engineering properties such as tensile strength,[12] ductility, and shear strength may be substantially different from those of the constituent materials.

Actiniums

[edit | edit source]
Actinium-225 medical radioisotope held in a v-vial at ORNL; note: the blue glow that originates from the ionization of surrounding air by alpha particles. Credit: Oak Ridge National Laboratory.{{free media}}
Uraninite ores have elevated concentrations of actinium. Credit: Robert Matthew Lavinsky.{{free media}}
Actinium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

AcOF, AcOCl, AcOBr exist.

Aluminums

[edit | edit source]
File:Aluminum1.jpg
Near the top center of this image is a gray reflective flake of native aluminum. Credit: Vasil Arnaudov.{{fairuse}}
File:Native aluminum in polished section 2.png
The bright silvery flakes are native aluminum in a polished section. Credit: Thomas Witzke / Abraxas-Verlag.{{fairuse}}
File:Aluminum flame.png
The image shows the color of aluminum in a natural gas burner. Credit: Alternative Fuels Laboratory/McGill University. {{fairuse}}
Aluminium spectrum is 400 nm - 700 nm Credit: McZusatz.{{free media}}

This flake was discovered, "During a field trip to the NW Rila Mountain in the early 1960s, one of us (V.A.) investigated the desilicated pegmatite apophysis and, from the phlogopite zone (Fig. 1c), collected a rock specimen with a protruding metallic flake visible to the naked eye (Fig. 2) [from which the above image was cropped]."[13]

The designation for native aluminum is Al0 as indicated in, "Here we present data for a unique Al0 flake protruding from the phlogopite matrix of a rock specimen collected from a desilicated pegmatite vein."[13]

Native aluminium metal is extremely rare and can only be found as a minor phase in low oxygen fugacity environments, such as the interiors of certain volcanoes.[14] Native aluminium has been reported in cold seeps in the northeastern continental slope of the South China Sea, where these deposits may have resulted from bacterial reduction of tetrahydroxoaluminate Al(OH)4.[15]

The second image of native aluminum is shown on the right of this section. The sample is from a mud volcano in the Caspian Sea near Baku, Azerbaidzhan.

The type locality for native aluminum is the Tolbachik volcano, Kamchatka, Russia.

Def. any "intermetallic compound of aluminium and a more electropositive element"[16] is called an aluminide.

The aluminides are those naturally occurring minerals with a high atomic % aluminum.

In the image on the right of a flake of native aluminum, the scale bar = 1 mm.

The typical alloying elements are copper, magnesium, manganese, silicon, tin and zinc.

Aluminium is the third most abundant element (after oxygen and silicon) in the Earth's crust, and the most abundant metal there. It makes up about 8% by mass of the crust, though it is less common in the mantle below.

As a mineral occurrence, aluminum is mostly an oxide.

Americiums

[edit | edit source]
A small disc of Am-241 is shown under the microscope. Credit: Bionerd.{{free media}}
Americium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Sorption of americium at trace levels has been detected on a clay mineral.[17]

Antimonies

[edit | edit source]
This is massive antimony with oxidation products from Arechuybo, Mexico. Credit: Aram Dulyan at the Natural History Museum, London.{{free media}}
The native antimony crystals, lustrous and nicely striated in part, range up to 0.5 cm in size. Credit: Robert Lavinsky.{{free media}}
The image shows the color of antimony in a natural gas burner. Credit: Herge.{{free media}}

.

Antimony spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Native antimony such as occurs in the rock on the upper right with its various oxidation products is crystalline in the hexagonal system.

The image on the left shows hexagonal crystals with metallic luster.

Argons

[edit | edit source]

Arsenics

[edit | edit source]
File:Native arsenic from silver vein.jpg
Native arsenic such as this specimen is found in silver ore veins. Credit: Amethyst Galleries, Inc.{{fairuse}}
This massive native arsenic with quartz and calcite is from Ste. Marie-aux-mines, Alsace, France. Credit: Aram Dulyan.{{free media}}
Ultrapure metallic arsenic is kept under argon, 1-2 grams. Credit: Hi-Res Images of Chemical Elements.{{free media}}
The image shows the color of arsenic in a natural gas burner. Credit: Herge.{{free media}}

.

Arsenic emission spectrum is for 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Native arsenic such as in the image on the right occurs in silver ore veins.

"The dominant group V source is arsenic, although antimony and phosphorous sources are not atypical."[18]

Allemontites

[edit | edit source]
Allemontite (with Pen for scale) is from the mineral collection of Brigham Young University Department of Geology, Provo, Utah. Credit: Andrew Silver, USGS.{{free media}}
Allemontite specimen is from Příbram, Central Bohemia Region, Bohemia (Böhmen; Boehmen), Czech Republic. Credit: Robert Lavinsky.{{free media}}

Allemontite is a native alloy of arsenic and antimony, with a composition of AsSb.[19]

The first example on the right is from the mineral collection of Brigham Young University Department of Geology, Provo, Utah.

The second on the left is from Příbram, Central Bohemia Region, Bohemia (Böhmen; Boehmen), Czech Republic.

As a natural source of arsenic, it has 50 at % arsenic.

Astatines

[edit | edit source]

Astatine iodide has the chemical formula AtI. Astatine bromide has the chemical formula AtBr. Astatine monochloride (AtCl) is made either by the direct combination of gas-phase astatine with chlorine or by the sequential addition of astatine and dichromate ion to an acidic chloride solution.

Uraninites

[edit | edit source]
File:Uraninite Astatine source.jpg
The rarest naturally occurring element on Earth is named Astatine and it occurs in uraninite as a uranium decay product. Credit: Fred E. Davis.{{fairuse}}
File:Uraninite crystal cluster.jpg
A cluster of seven crystals, four are visible, in the photo of uraninite, with a yellow uranophane coating. Credit: Fred E. Davis.{{fairuse}}

All of the known isotopes of astatine are very short-lived. Astatine occurs naturally in minerals such as uraninite as a decay product of uranium.

Bariums

[edit | edit source]
Pure barium metal in a protective argon gas atmosphere. Credit: Matthias Zepper.{{free media}}
File:Barium flame.png
The image shows the color of barium in a natural gas burner. Credit: Ernest Z. {{fairuse}}
This is a pressure-temperature phase diagram for barium. Credit: David A. Young, ERDA.{{free media}}
Barium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Barium is bcc (α-Ba) at room temperature as the phase diagram on the left indicates. It does change to an hcp structure at high pressures and temperatures.

Native barium is not known to occur on the surface of the Earth.

Berkeliums

[edit | edit source]
The geological situation in Gabon leading to natural nuclear fission reactors is described
1. Nuclear reactor zones
2. Sandstone
3. Uranium ore layer
4. Granite. Credit: MesserWoland.{{free media}}
Berkelium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Like americium and curium it "is possible that some berkelium and other transuranic elements were created in the natural nuclear reactor in Oklo, Gabon."[20]

A natural nuclear fission reactor is a uranium mineral deposit where self-sustaining nuclear chain reactions have occurred. This can be examined by analysis of isotope ratios. The existence of this phenomenon was discovered in 1972 at Oklo in Gabon, Africa. Oklo is the only known location for this in the world and consists of 16 sites at which self-sustaining nuclear fission reactions took place approximately 1.7 billion years ago, and ran for a few hundred thousand years, averaging 100 kW of thermal power during that time.[21][22]

Berylliums

[edit | edit source]
This spectrograph shows the visual spectral lines of beryllium. Credit: Penyulap.{{free media}}
Beryllium is > 99 % pure, crystalline big fragment > 140 g. Credit: Alchemist-hp.{{free media}}
File:Beryllium-Chromium phase diagram.png
This is a beryllium-chromium phase diagram. Credit: M. Venkatraman and J.P. Neumann.{{fairuse}}

Beryllium has at least six emission/absorption lines across the red.

The emission and absorption spectra for beryllium contain lines in the blue.

Beryllium occurs in a hexgonal close-packed (hcp) crystal structure at room temperature (α-Be).

As indicated in the phase diagram on the left beryllium occurs as (β-Be) which is bcc at higher temperatures up to melting.

Native beryllium is not known to occur on the surface of the Earth, but may eventually be found among beryllium-bearing minerals in small amounts.

Beryllium copper (BeCu), also known as copper beryllium (CuBe), beryllium bronze and spring copper, is a copper alloy with 0.5–3% beryllium,[23] but can contain other elements as well. Beryllium can be alloyed with nickel and aluminum.[24]

Bismuths

[edit | edit source]
A rich small mini of the native element bismuth is from from China. Credit: Robert M. Lavinsky.{{free media}}
Bismuth spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Bismuth does occur on Earth as native bismuth exampled on the right.

"Strong absorption lines due to Bi II have been found in the Hg—Mn star HR7775 (HD193452) in high-resolution spectra obtained with the IUE."[25]

"Detailed examination of the optical spectrum at high resolution1 showed that it is one of the most extreme stars of the ‘cool’ Hg—Mn group, with strong enhancements of Hg, Pt, Sr, Y, and Ga; the last of these is confirmed2 by the very strong Ga II resonance line at 1,414 Å. Four-colour Strömgren photometry of HR7775 (ref. 3), interpreted with the aid of the model atmosphere calibrations by Relyea and Kurucz4, gives Teff = 10,800 K, log g = 4.2, while the Hβ index gives log g = 4.0 according to the calibration of Schmidt5."[25]

Bohriums

[edit | edit source]

Chemistry experiments have confirmed that bohrium behaves as the heavier homologue to rhenium in group 7. The chemical properties of bohrium are characterized only partly, but they compare well with the chemistry of the other group 7 elements.

Borons

[edit | edit source]
These polycrystalline chunks are rhombohedral β-boron, net 25.5 grams. Credit: James L Marshall.{{free media}}
Phase diagram is elemental boron (colors represent actuals). Credit: Fulvio314.{{free media}}
This spectrograph shows the visual spectral lines of boron. Credit: Penyulap.{{free media}}
Crystalline boron, which is shown here, is nearly as hard as diamond (9.5 on Mohs scale, diamond has 10). Credit: Hi-Res Images of Chemical Elements.{{free media}}
The image shows the color of boron in a natural gas burner. Credit: Herge.{{free media}}

.

Boron is synthesized entirely by cosmic ray spallation and supernovae and not by stellar nucleosynthesis, so it is a low-abundance element in the Solar System and in the Earth's crust.[26] It constitutes about 0.001 percent by weight of Earth's crust.[27] It is concentrated on Earth by the water-solubility of its more common naturally occurring compounds, the borate mineral such as borax and kernite.

Elemental boron is a metalloid that is found in small amounts in meteoroids but chemically uncombined boron is not otherwise found naturally on Earth.

The "presence in ... cosmic radiation [is] of a much greater proportion of "secondary" nuclei, such as lithium, beryllium and boron, than is found generally in the universe."[28]

Qingsongites

[edit | edit source]

Qingsongite is a rare boron nitride (BN) mineral with cubic crystalline form first described in 2009 for an occurrence as minute inclusions within chromite deposits in the Luobusa ophiolite in the Shannan Prefecture, Tibet Autonomous Region, China.[29] It was recognized as a mineral in August 2013 by the International Mineralogical Association named after Chinese geologist Qingsong Fang (1939–2010).[29]

Qingsongite is the only known boron mineral that is formed deep in the Earth's mantle.[30] Associated minerals or phases include osbornite (titanium nitride), coesite, kyanite and amorphous carbon.[31]

Wurtzite BN

[edit | edit source]

Only small amounts of the wurtzite form of boron nitride (w-BN) exist in nature as a mineral.[32]

Bromines

[edit | edit source]

Cadmiums

[edit | edit source]
Cadmium metal bowl is shown. Credit: Halfdan.{{free media}}
A crystal cadmium bar, purity 99.999 %, made by the flux process, as well as a 1 cm3 cadmium cube for comparison. Credit: Alchemist-hp.{{free media}}
Cadmium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The only cadmium mineral of importance, greenockite (CdS), is nearly always associated with sphalerite (ZnS).

Cadmium is used in the control rods of nuclear reactors, acting as a very effective neutron poison to control neutron flux in nuclear fission.[33] When cadmium rods are inserted in the core of a nuclear reactor, cadmium absorbs neutrons, preventing them from creating additional fission events, thus controlling the amount of reactivity. The pressurized water reactor designed by Westinghouse Electric Company uses an alloy consisting of 80% silver, 15% indium, and 5% cadmium.[33]

In Polyvinyl chloride (PVC), cadmium was used as heat, light, and weathering stabilizers.[33][34] Cadmium is used in many kinds of solder and bearing alloys, because it has a low coefficient of friction and fatigue resistance.[33] It is also found in some of the lowest-melting alloys, such as Wood's metal.[35]

Caesiums

[edit | edit source]
High pure cesium crystals show dendritic morphology. Credit: Dnn87.{{free media}}
Temperature-pressure diagram for caesium, formerly known as "cesium". Credit: David A. Young, ERDA.{{free media}}
File:Caesium flame.png
The image shows the color of caesium in a natural gas burner. Credit: Ernest Z. {{fairuse}}
Caesium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

As the temperature-pressure diagram on the left shows, caesium (formerly cesium) is bcc (α-Cs) from room temperature up to melting.

Native caesium does not appear to occur on the surface of the Earth or the Moon.

Calciums

[edit | edit source]
Pure calcium metal is shown in a protective argon atmosphere. Credit: Matthias Zepper.{{free media}}
Calcium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
The image shows the color of calcium in a natural gas burner. Credit: Herge.{{free media}}

.

Spectrum of Ca is at 600/pmm. Credit: teravolt.{{free media}}
File:Aluminum-Calcium phase diagram.png
This is an Aluminum-Calcium phase diagram. Credit: The ESA IMPRESS Team.{{fairuse}}

Calcium has a face-centered cubic (fcc) crystal structure at room temperature.

As shown in the phase diagram on the left, it does not change structure up to melting.

Native calcium is not known to occur on the surface of the Earth.

Californiums

[edit | edit source]
A disc of californium metal (249Cf, 10 mg). Credit: United States Department of Energy.{{free media}}
Californium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

It forms alloys with lanthanide metals.[36]

The element has two crystalline forms at standard atmospheric pressure: a double-hexagonal close-packed form dubbed alpha (α) and a face-centered cubic form designated beta (β). A double hexagonal close-packed (dhcp) unit cell consists of two hexagonal close-packed structures that share a common hexagonal plane, giving dhcp an ABACABAC sequence.[37] The α form exists below 600–800 °C with a density of 15.10 g/cm3 and the β form exists above 600–800 °C with a density of 8.74 g/cm3.[36] At 48 GPascal of pressure the β form changes into an orthorhombic crystal system due to delocalization of the atom's 5f electrons, which frees them to bond.[36] The three lower-mass transplutonium elements—americium, curium, and berkelium—require much less pressure to delocalize their 5f electrons.[36]

Carbons

[edit | edit source]
The spectrum shows the lines in the visible due to emission from elemental carbon. Credit:Teravolt.{{free media}}
The image shows the color of carbon in a natural gas burner. Credit: Arthur Jan Fijałkowski. {{free media}}

.

Diamond and graphite are shown side by side, where the scale is based on a rough approximation. Credit: Robert Lavinsky.{{free media}}

Carbonides are naturally occurring minerals composed of 50 atomic percent, or more, carbon. Carbonide-like minerals with greater than 25 at % carbon are also included. This separates carbon containing minerals from carbonates which are at most 25 at % carbon.

Carbon has an emission line in plasmas at 529.053 nm from C VI.[38]

Ceriums

[edit | edit source]
Ultrapure Cerium metal is shown under argon, 1.5 grams. Credit: Jurii.{{free media}}
Cerium spectrum is shown for 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The pyrophoric alloy known as "mischmetal" is composed of 50% cerium, 25% lanthanum, and the remainder being the other lanthanides, that is used widely for lighter flints.[39] Usually iron is added to form the alloy ferrocerium (a synthetic pyrophoric alloy of "mischmetal": cerium, lanthanum, neodymium, other trace lanthanides and some iron – about 95% lanthanides and 5% iron hardened by blending in oxides of iron and / or magnesium).[40]

Cerium is used as alloying element in aluminum to create castable eutectic aluminum alloys with 6–16 wt.% Ce, to which Mg and/or Si can be further added, which have excellent high temperature strength and are suitable for automotive applications e.g. in cylinder heads.[41] Other alloys of cerium include Pu-Ce and Pu-Ce-Co plutonium alloys, which have been used as nuclear fuel.

Chlorines

[edit | edit source]

Chromiums

[edit | edit source]
File:Native chromium.jpg
This is a native chromium nugget. Credit: Neal Ekengren.{{fairuse}}
File:Fe-Cr Phase Diagram.gif
Fe-Cr phase diagram shows which phases are to be expected at equilibrium for different combinations of chromium content and temperature. Credit: Computational Thermodynamics Inc.{{fairuse}}
A hunk of chromium metal is shown. Credit: W. Oelen.{{free media}}
The image shows the color of chromium in an oxy-hydrogen torch. Credit: NSEasternShoreChemist.{{free media}}

.

This is a visible emission-line spectrum for chromium over the range: 400-700 nm. Credit: McZusatz.{{free media}}

Native chromium such as the nugget in the image on the right is very rare. It is also a hard mineral, probably because of an oxide coating giving it a slight bluish cast.

"An unusual mineral association (diamond, SiC, graphite, native chromium, Ni-Fe alloy, Cr2+-bearing chromite), indicating a high-pressure, reducing environment, occurs in both the peridotites and chromitites."[42]

As the phase diagram for the Fe-Cr system on the left shows, chromium is bcc from 600°C on up to melting. Chromium is also bcc at room temperature and pressure.

Cobalts

[edit | edit source]
File:Native cobalt from Moon.jpg
This is a scanning electron micrograph of native cobalt from the Luna 24 landing site, Mare Crisium, The Moon. Credit: Pavel M. Kartashov, Hudson Institute of Mineralogy.{{fairuse}}
Pure (99.9 %) cobalt chips, electrolytically refined, and a high purity (99.8 % = 2N8) 1 cm3 cobalt cube for comparison is shown. Credit: Alchemist-hp.{{free media}}
The image shows the color of cobalt in an oxy-hydrogen torch. Credit: NSEasternShoreChemist.{{free media}}

.

Cobalt emission spectrum is from 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Cobalt has a hexagonal close-packed structure (hcp) until about 450°C when a fcc structure begins to appear.

On the right is a scanning electron micrograph of native cobalt from the Luna 24 landing site, Mare Crisium, The Moon.

Coperniciums

[edit | edit source]

Very few properties of copernicium or its compounds have been measured; this is due to its extremely limited and expensive production[43] and the fact that copernicium (and its parents) decays very quickly.

Coppers

[edit | edit source]
A large, sculptural specimen of penny-bright copper is from Arizona. Credit: Robert Lavinsky.{{free media}}
Copper emission spectrum is for 400 nm - 700 nm. Credit: McZusatz.{{free media}}
File:Landsat Thematic Mapper Manitouwadge.jpg
This is a Landsat Thematic Mapper image with overlain geological structures. Credit: I.M. Kettles, A.N. Rencz, and S.D. Bauke.{{fairuse}}
File:Copper (I) blue flame.png
The image shows the color of copper (I) in a natural gas burner. Credit: Anne Marie Helmenstine. {{fairuse}}
The image shows the color of copper (II) non-halide in a natural gas burner. Credit: Søren Wedel Nielsen.{{free media}}

.

File:Cu+2 (CuCl2)-Blue.jpg
The image shows the color of copper (II) halide in a natural gas burner. Credit: AmazingRust. {{fairuse}}

The most advantageous form for copper is native copper.

On the right is a large, sculptural specimen of penny-bright copper from Arizona.

"Approximately five million tonnes were mined from native copper deposits in Michigan. Copper masses from the Michigan deposits were transported by the Pleistocene glaciers. Areas on the copper surfaces which appear to represent glacial abrasion show minimal corrosion."[44]

"A group of pixel areas north of Lake Superior [in the Landsat image on the right] take the form of a linear band which lies along the northern edge of the Port Coldwell Complex (D). [...] there are numerous Cu showings along the northern edge of the Port Coldwell complex (Ontario Division of Mines, 1971)."[45]

Classification of copper and its alloys
Family Principal alloying element Composition range wt % Other elements
Copper alloys, brass Zinc (Zn) 30% Zn 0.02–0.15% As, 1.7–2.8% Pb, P, Al, Mn, 0.05% iron, and Si
Phosphor bronze Tin (Sn) 0.5–11% Sn, 0.01–0.35% P 0.5–3.0% Pb
Aluminium bronzes Aluminium (Al) 5% to 11% aluminium iron, nickel, manganese, zinc, silicon, arsenic
Silicon bronzes Silicon (Si) <6%Si, 92.5% Cu-7.5% Si Al, Zn, Ti, Fe
Cupronickel, nickel silvers Nickel (Ni) 60-90% Cu, 9-32% Ni, ≥52% Ni (Monel) 0.4-2.3% Fe, 1-2.5% Mn

Curiums

[edit | edit source]
Curium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

"Curium is a radioactive transuranic element that has only been produced in nuclear reactors. It is possible that some curium and other transuranic elements were created in the natural nuclear reactor in Oklo, Gabon."[46]

Darmstadiums

[edit | edit source]

The only known darmstadtium isotope with a half-life long enough for chemical research is 281Ds, which would have to be produced as the granddaughter of 289Fl.[47]

Dubniums

[edit | edit source]

Dubnium was processed in nitric and hydrofluoric acid solution, at concentrations where niobium forms NbOF
4
and tantalum forms TaF
6
, where dubnium's behavior was close to that of niobium but not tantalum; it was thus deduced that dubnium formed DbOF
4
, it was concluded that dubnium often behaved like niobium, sometimes like protactinium, but rarely like tantalum.[48]

Dysprosiums

[edit | edit source]
Dysprosium metal chips are shown. Credit: Materialscientist.{{free media}}
Dysprosium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
First order and partial second order spectra of a 12V dysprosium spark taken with a diffraction grating and a cellphone camera. Credit: Umop503.{{free media}}

"The magnetic and structural properties of the neodymium-dysprosium alloy system have been measured over the entire composition range."[49]

Einsteiniums

[edit | edit source]
Quartz vial (9 mm diameter) contains ~300 micrograms of Es-253 solid. Credit: Haire, R. G., US Department of Energy.{{free media}}
Einsteinium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Einsteinium is a soft, silvery, paramagnetic metal with chemistry typical of the late actinides, with a preponderance of the +3 oxidation state; the +2 oxidation state is also accessible, especially in solids.

Erbiums

[edit | edit source]
Erbium is a silvery-white solid metal. Credit: Tomihahndorf.{{free media}}
Erbium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
12V spark spectrum of erbium taken through a diffraction grating with a cellphone camera. Credit: Umop503.{{free media}}

When added to vanadium as an alloy, erbium lowers hardness and improves workability.[50] An erbium-nickel alloy Er3Ni has an unusually high specific heat capacity at liquid-helium temperatures.

"Along with uranium, zinc, iron ore, copper and gold, Greenland’s ancient rocks also harbor large quantities of those minerals known as “rare earth,” among them lanthanum, cerium, neodymium, praesodymium, terbium and yttrium."[51]

Europiums

[edit | edit source]
Weakly oxidized europium, hence slightly yellowish, 1.5 grams, large piece 0.6 x 1.6 cm are shown. Credit: Jurii.{{free media}}
Europium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Spark spectrum of europium metal taken using a 12V circuit, diffraction grating, and phone camera. Note this is the spectrum of the pure element, not the spectrum it makes as a dopant in phosphors. Credit: Umop503.{{free media}}

The "aluminum−boron−europium ternary alloy fuels with boron content of 1.5∼4.85 wt. % and europium content of 3 wt. % were prepared."[52]

Fermiums

[edit | edit source]
A fermium-ytterbium alloy is used for measuring the enthalpy of vaporization of fermium metal. Credit: Ben E. Lewis.{{free media}}

In the image at the right a fermium-ytterbium alloy is shown.

Flevoriums

[edit | edit source]

About 90 flerovium atoms have been seen: 58 were synthesized directly; the rest were from radioactive decay of heavier elements.

Fluorines

[edit | edit source]

Franciums

[edit | edit source]
This rock probably contains on the order of a few atoms of francium at any one time, as part of the complex decay chains of the thorium and uranium that make up a much larger fraction of the sample. Credit: Theodore Gray.{{free media}}
This sample of uraninite contains about 100,000 atoms (3.3×1020 g) of francium-223 at any given time.[53] Credit: Farhan.{{free media}}

Francium is bcc at room temperature. Outside the laboratory, francium is extremely rare, with trace amounts found in uranium and thorium ores, where the isotope francium-223 continually forms and decays.

Francium chloride has been studied as a pathway to separate francium from other elements, by using the high vapour pressure of the compound, although francium fluoride would have a higher vapour pressure.[54]

Gadoliniums

[edit | edit source]
Gadolinium is a silvery-white metal when oxidation is removed. Credit: Unknown author.{{free media}}
Gadolinium is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Spectrum of gadolinium from a 12V spark, taken through a diffraction grating with a cellphone camera. The full first order spectrum is visible, and most of the second order; a spectrum of a reflection is also visible, showing the unusual color combination this element emits. Credit: Umop503.{{free media}}

Gadolinium metal is only slightly malleable and is a ductile rare-earth element.

Gadolinium demonstrates a magnetocaloric effect whereby its temperature increases when it enters a magnetic field and decreases when it leaves the magnetic field. The temperature is lowered to 5 °C (41 °F) for the gadolinium alloy Gd85Er15, and this effect is considerably stronger for the alloy Gd5(Si2Ge2), but at a much lower temperature (<85 K (−188.2 °C; −306.7 °F)).[55] A significant magnetocaloric effect is observed at higher temperatures, up to about 300 K, in the compounds Gd5(SixGe1−x)4.[56]

Galliums

[edit | edit source]
File:Gallite from Namibia.jpg
This is an example of gallite from Namibia. Credit: Hudson Institute of Mineralogy.{{fairuse}}
This is a liquid drop of gallium. Credit: RTC.{{free media}}
This solid gallium is fresh and after some time (2 months) at room temperature. Credit: Hi-Res Images of Chemical Elements.{{free media}}
Gallium emission spectrum is for 400 nm - 700 nm. Credit: McZusatz.{{free media}}

While native gallium would be the best source of gallium, it apparently does not occur on Earth. The image on the right is a drop of liquid gallium.

Gallium "enrichments are observed in the deep waters of the Norwegian Sea and Iceland Basin."[57]

"If northern deep water formation occurs at lower latitudes during glacial periods, the amount of sediment resuspension in the formation areas is likely to be affected with concomitant effects on the trace element content of newly formed northern-source deep waters."[57]

"At higher growth temperatures(>600'C) the lifetimes of the alkyl-gallium species are much shorter and the growth front dynamics should begin to look more like MBE since atomic gallium will be the dominant group III surface species."[58]

Gallite (CuGaS2) is 25 at % gallium.

Germaniums

[edit | edit source]
File:Germanite from Tsumeb, Namibia.jpg
This sample of germanite is displayed in the Smithsonian Museum of Natural History. Credit: R Nave.{{fairuse}}
This chunk is 12 grams of polycrystalline germanium, 2 x 3 cm. Credit: Jurii.{{free media}}
Germanium emission spectrum is for 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The sample of germanite on the right has a composition of Cu26Fe4Ge4S32. Generally, germanite has a composition closer to Cu3(Ge, Ga, Fe, Zn) (S,As)4.[19] "This sample also contains tennantite."[59]

Golds

[edit | edit source]
This very elaborate, 3-dimensional cluster of gold shows complex and minute crystallization patterns, and is overall hackly in texture. Credit: Robert Lavinsky.{{free media}}
This gold nugget (placer gold) is 9.5 mm across at its widest. Credit: James St. John.{{free media}}
Gold spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Gold (Au) is the most prestigious metal known, but it's not the most valuable. Gold is the only metal that has a deep, rich, metallic yellow color. Almost all other metals are silvery-colored. Gold is very rare in crustal rocks - it averages about 5 ppb (parts per billion). Where gold has been concentrated, it occurs as wires, dendritic crystals, twisted sheets, octahedral crystals, and variably-shaped nuggets. It most commonly occurs in hydrothermal quartz veins, disseminated in some contact- & hydrothermal-metamorphic rocks, and in placer deposits. Placers are concentrations of heavy minerals in stream gravels or in cracks on bedrock-floored streams. Gold has a high specific gravity (about 19), so it easily accumulates in placer deposits. Its high density allows prospectors to readily collect placer gold by panning.

Hafniums

[edit | edit source]
File:Fe-Hf phase diagram.gif
This is an iron-hafnium phase diagram. Credit: H. Okamoto.{{fairuse}}
A sample of a 1,7kg Hafnium crystal bar, made by van Arkel-de Boer process. Credit: Alchemist-hp.{{free media}}
Hafnium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Note in the iron-hafnium phase diagram on the left that hafnium occurs in two phases: hcp (α-Hf) at lower temperatures and bcc (β-Hf) at higher temperatures up to melting.

Hafnium is used in alloys with iron, titanium, niobium, tantalum, and other metals, for example the main engine of the Apollo Lunar Modules, is C103 which consists of 89% niobium, 10% hafnium and 1% titanium.[60]

Small additions of hafnium increase the adherence of protective oxide scales on nickel-based alloys, improving thereby the corrosion resistance especially under cyclic temperature conditions that tend to break oxide scales by inducing thermal stresses between the bulk material and the oxide layer.[61][62][63]

Hassiums

[edit | edit source]

Hassium behaves as the heavier homologue to osmium, reacting readily with oxygen to form a volatile tetroxide.

Heliums

[edit | edit source]

Holmiums

[edit | edit source]
Ultrapure holmium metal of 17 grams is shown. Credit: Unknown author.{{free media}}
Holmium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Spectrum of holmium from a 12V spark, taken through a diffraction grating with a cellphone camera. Credit: Umop503.{{free media}}

"The constant of the alloy-formation rate for HoNi
2
, which was obtained in [22] at 1023 K in the LiCl–KCl eutectic melt, is 0.36 kg/m2 h0.5."[64]

Hydrogens

[edit | edit source]

Indiums

[edit | edit source]
File:Indium R060771 Eastern Transbaikal, Russia.jpg
These pieces of native indium are from Eastern Transbaikal, Russia. Credit: Michael Scott.{{fairuse}}
Indium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Indium is an ingredient in the gallium–indium–tin alloy galinstan, which is liquid at room temperature and replaces mercury in some thermometers.[65] Other alloys of indium with bismuth, cadmium, lead, and tin, which have higher but still low melting points (between 50 and 100 °C), are used in fire sprinkler systems and heat regulators.[39]

"Indium minerals are very rare ; only 7 species have been defined so far : roquesite, CuInS2 (Picot & Pierrot, 1963) ; indite, FeIn2S4, and dzhalindite, In(OH)3 (Genkin & Murav'eva, 1963) ; sakuraiite, (Cu,Fe,Zn)3(In,Sn)S4 (Kato, 1965) ; native indium (Ivanov, 1966b) ; yixunite, PtIn (Yu Tsu-Hsiang et al., 1976) ; petrukite, (Cu,Fe,Zn,Ag)3(Sn,In)S4 (Kissin & Owens, 1989)."[66]

On the right are microprobe fragments of native indium from Eastern Transbaikal, Russia. The electron microprobe confirms that indium is the only component of the metallic phase.

Iodines

[edit | edit source]

Iridiums

[edit | edit source]
File:Cubic native iridium crystal.png
The image is of a cubic crystal of native iridium. Credit: Norman King.{{fairuse}}
Pieces are pure iridium, 1 gram. Credit: Unknown author.{{free media}}
Iridium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Native iridium such as the small cubic crystal shown in the image on the right is rare.

An alloy of iridium with ruthenium in thermocouples allowed for the measurement of high temperatures in air up to 2,000 °C (3,630 °F).[67]

Iridium is found in nature as an uncombined element or in natural alloys; especially the iridium–osmium alloys, osmiridium (osmium-rich), and iridosmium (iridium-rich).[68]

Irons

[edit | edit source]
File:Native iron.jpg
This piece from the Khungtukun Massif, Malaya Romanikha River, Khatanga, Taimyr Peninsula, Taymyrskiy Autonomous Okrug, Eastern-Siberian Region, Russia, displays inclusions of native iron. Credit: Hudson Institute of Mineralogy.{{fairuse}}
This diagram shows most of the phases of pure iron. Credit: Daniele Pugliesi and Materialscientist.{{free media}}
The image shows the color of iron (III) orange in a thermite reaction. Credit: Denver & Rio Grande. {{free media}}

.

File:Epsilon iron unary phase diagram.jpg
The unary (temperature-pressure) iron phase diagram shows the epsilon phase of iron at extremely high pressure. Credit: His Manliness.{{free media}}
An uncommon slabbed and polished specimen of lustrous, metallic, elemental native iron in basalt from Germany. Credit: Robert Lavinsky.{{free media}}
Iron is a silvery, relatively reactive metal, which is very abundant and is used for multiple purposes. Credit: Hi-Res Images of Chemical Elements.{{free media}}
Iron spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The polished piece on the top right displays inclusions of native iron.

Iron occurs in several allotropes from α-Fe which has a body-centered cubic structure (bcc) at room temperature up to 910°C, γ-Fe which has a face-centered cubic (fcc) structure from 910°C to 1394°C, and δ-Fe (bcc) from 1394°C to 1538°C. Hexagonal close-packed (hcp) iron occurs at high pressures and temperatures as ε-Fe.

Austenite, also known as gamma-phase iron (γ-Fe), is a metallic, non-magnetic allotrope of iron or a solid solution of iron, with an alloying element.[69] In plain-carbon steel, austenite exists above the critical eutectoid temperature of 1000 K (727 °C); other alloys of steel have different eutectoid temperatures. The austenite allotrope is named after Sir William Chandler Roberts-Austen (1843–1902);[70] it exists at room temperature in some stainless steels due to the presence of nickel stabilizing the austenite at lower temperatures.

Meteoritic irons

[edit | edit source]

Iron hydrides

[edit | edit source]

"Carroll and McCormack (1972) in Dublin reported complex spectra in the blue and green wavelength regions of both FeH and FeD".[71]

"Carroll et al. (1976) detected a number of coincidences between laboratory lines of FeH and weak unidentified solar lines, again in the blue and green wavelength region, in addition to the infrared."[72]

Kryptons

[edit | edit source]

Lanthanums

[edit | edit source]
1 cm big piece is pure lanthanum. Credit: Jurii.{{free media}}
Lanthanum spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Hydrogen sponge alloys can contain lanthanum and are capable of storing up to 400 times their own volume of hydrogen gas in a reversible adsorption process, where heat energy is released every time they do so; therefore these alloys have possibilities in energy conservation systems.[73][74]

Mischmetal, a pyrophoric alloy used in lighter flints, contains 25% to 45% lanthanum.[50]

Lawrenciums

[edit | edit source]

The first ionization energy of lawrencium was measured, using the isotope 256Lr.[75] The measured value, 4.96+0.08
−0.07
electronvolt (eV)
, agreed very well with the relativistic theoretical prediction of 4.963(15) eV, and also provided a first step into measuring the first ionization energies of the transactinides.[75] This value is the lowest among all the lanthanides and actinides, and supports the s2p configuration as the 7p1/2 electron is expected to be only weakly bound. This suggests that lutetium and lawrencium behave similarly to the d-block elements (and hence being the true heavier congeners of scandium and yttrium, instead of lanthanum and actinium). Although some alkali metal-like behaviour has been predicted,[76] adsorption experiments suggest that lawrencium is trivalent like scandium and yttrium, not monovalent like the alkali metals.[77]

Leads

[edit | edit source]
This is a piece of native lead. Credit: Rob Lavinsky.{{free media}}
A piece of lead, cut through, is silvery for a short time, before the surface oxidizes. Credit: Hi-Res Images of Chemical Elements.{{free media}}
The image shows the color of lead in a natural gas burner. Credit: Herge.{{free media}}

.

Lead spectrum is for 400 nm - 700 nm. Credit: McZusatz.{{free media}}

"Diamond cubic structures with lattice parameters around the lattice parameter of silicon exists both in thin lead and tin films, and in massive lead and tin, freshly solidified in vacuum of ≈5 x 10-6 Torr. Experimental evidence for almost identical structures of at least three oxide types is presented, demonstrating that lead and tin behave like silicon not only in the initial stages of crystallization, but also in the initial stages of oxidation."[78]

The piece of native lead on the right shows a relatively sharp, and well-formed cuboctahedron of Lead at the top of the specimen, which is associated with elongated crystals on the base and back.

Its source locality is Långban, Filipstad, Värmland, Sweden.

A fresh surface of high purity lead on the left is silvery in appearance.

Lead tin telluride, PbSnTe or Pb1−xSnxTe, is a ternary alloy of lead, tin and tellurium, generally made by alloying either tin into lead telluride or lead into tin Telluride.

In genuine Ashtadhatu, all eight metals (Au, Ag, Cu, Pb, Zn, Sn, Fe and Sb or Hg) are in equal proportion (12.5% each).[79][80][81]

Sn, Pb, Cu, As, Sb are used to make Babbitt alloys.[82] "The inner parts of the boxes are to be lined with any of the harder kinds of composition known under the names of britannia metal or pewter, of which block tin is the basis. An excellent compound for this purpose I have prepared by taking about 50 parts of tin, five of antimony, and one of copper, but I do not intend to confine myself to this particular composition."[82]

Lithiums

[edit | edit source]
The image shows the color of lithium in a natural gas burner. Credit: Herge.{{free media}}

.

This spectrograph shows the visual spectral lines of lithium. Credit: T c951.{{free media}}
Lithium remains chemically inert when immersed in oil. Credit: Tomihahndorf.{{free media}}

"[T]he standard solar models have enjoyed tremendous success recently in terms of agreement between the predicted outer structure and the results from helioseismology[, but] some observed properties of the Sun still defy explanation, such as the degree of Li depletion" [the "solar Li abundance is roughly a factor of 200 below the meteoritic abundance"].[83]

Livermoriums

[edit | edit source]

Although generated by heavy ion bombardment, the short-lived radioisotopes are not known to occur naturally on the surface of the Earth.

Lutetiums

[edit | edit source]
Lutetium, sublimed-dendritic, high purity 99.995 % Lu/TREM, as well as an argon arc remelted 1 cm3 lutetium (99,9 %) cube for comparison. Credit: Alchemist-hp.{{free media}}
Lutetium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Scandium, yttrium, and lutetium tend to occur together with the other lanthanides (except short-lived promethium) in the Earth's crust, and are often harder to extract from their ores.

Magnesiums

[edit | edit source]
This image shows the optical emission lines of magnesium. Credit: teravolt.{{free media}}
Crystalized magnesium is shown. Credit: Mark Fergus, CSIRO.{{free media}}
File:Magnesium-rich end with iron.png
This portion of the iron-magnesium phase diagram is concentrated on the magnesium-rich end. Credit: A.A. Nayeb-Hashemi, J.B. Clark and L.J. Swartzendruber.{{fairuse}}
File:Mg-flame.jpg
The image shows the color of magnesium in a natural gas burner. Credit: KB. {{fairuse}}

Magnesium has a hcp structure from room temperature up to melting. No other phases occur as is shown in the magnesium-end of the iron-magnesium phase diagram on the left.

Native magnesium is unlikely to occur on the surface of the Earth and is not known to occur.

Magnesium (Mg I) has an absorption band at 416.727±2.9 nm with an excitation potential of 4.33 eV.[84]

Magnesium (Mg II) has an absorption band at 439.059±6.6 nm with an excitation potential of 9.96 eV.[84]

Manganeses

[edit | edit source]
File:Fe-Mn Phase Diagram.gif
The pure manganese end member of this Fe-Mn phase diagram shows the higher temperature Mn phases. Credit: A. Rabinkin.
Two pieces of manganese metal are shown. Credit: W. Oelen.{{free media}}
The image shows the color of manganese (II) in an oxy-hydrogen torch. Credit: NSEasternShoreChemist.{{free media}}

.

Manganese emission spectrum is from 400 nm - 700 nm Credit: McZusatz.{{free media}}

If native manganese occurs on Earth or nearby Solar System bodies, it likely occurs as bcc α-Mn.

"Beta manganese has a cubic crystal structure with space group P4132 [1]. The unit cell contains 20 atoms, divided between two non-equivalent sites."[85]

"The structures of γ- and δ-manganese are found to be face-centred cubic and body-centred cubic respectively."[86]

Manganese (Mn I) has two absorption bands at 403.449±1.4 nm and 405.554±0.8 nm, where the second has an excitation potential of 2.13 eV.[84]

Manganese (Mn II) has an absorption band at 420.638±0.8 nm with an excitation potential of 5.37 eV.[84]

Meitneriums

[edit | edit source]

Meitnerium is the seventh member of the 6d series of transition metals, and should be much like the platinum group metals.[87] Calculations on its ionization potentials and atomic radius and ionic radii are similar to that of its lighter homologue iridium, thus implying that meitnerium's basic properties will resemble those of the other group 9 elements, cobalt, rhodium, and iridium.[77]

Mendeleviums

[edit | edit source]

Seventeen isotopes of mendelevium are known, with mass numbers from 244 to 260; all are radioactive.[88] Additionally, five nuclear isomers are known: 245mMd, 247mMd, 249mMd, 254mMd, and 258mMd.[89][90] Of these, the longest-lived isotope is 258Md with a half-life of 51.5 days, and the longest-lived isomer is 258mMd with a half-life of 58.0 minutes.[89][90] Nevertheless, the shorter-lived 256Md (half-life 1.17 hours) is more often used in chemical experimentation because it can be produced in larger quantities from alpha particle irradiation of einsteinium.[88]

Mercuries

[edit | edit source]

Metalloids

[edit | edit source]
This massive native arsenic with quartz and calcite is from Ste. Marie-aux-mines, Alsace, France. Credit: Aram Dulyan.{{free media}}

Metalloids are elements whose properties are intermediate between metals and solid nonmetals or semiconductors.

A variety of elements are often considered metalloids:

  1. boron, considered here in Borons,
  2. aluminum, a face-centered cubic metal, considered in Aluminums,
  3. silicon, here in Silicons,
  4. gallium, here in Galliums,
  5. germanium, here in Germaniums,
  6. arsenic, here in Arsenics,
  7. selenium, here in Seleniums, also included in the chalcogens,
  8. indium, here in Indiums,
  9. tin, here in Tins,
  10. antimony, here in Antimonies,
  11. tellurium, here in Telluriums, also included in the chalcogens,
  12. polonium, here in Poloniums, and
  13. astatine, here in Astatines, with the halogens.

Moscoviums

[edit | edit source]

The moscovium isotopes 288Mc, 289Mc, and 290Mc may be chemically investigated with current methods, although their short half-lives would make this challenging.[91] Moscovium is the heaviest element that has known isotopes that are long-lived enough for chemical experimentation.[47]

Molybdenums

[edit | edit source]
File:Native molybdenum Luna 24 landing site.jpg
This is a scanning electron micrograph of native molybdenum particles in lunar regolith. Credit: A. V. Mokhov and P. M. Kartashov.{{fairuse}}
Molybdenum spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
File:Fe-Mo Phase Diagram.gif
This is a calculated iron-molybdenum phase diagram. Credit: Computational Thermodynamics Inc.{{fairuse}}

The electron micrograph on the right shows a couple of pieces of native molybdenum found in lunar regolith at the Luna 24 landing site after transport back to Earth and analysis.

The phase diagram for the iron-molybdenum system demonstrates that molybdenum is bcc (α-Mo) for its intermediate and higher temperatures. It's also bcc at room temperature.

Molybdenum can withstand extreme temperatures without significantly expanding or softening, making it useful in environments of intense heat.[53][92]

Neodymiums

[edit | edit source]
Ultrapure neodymium is shown under argon, 5 grams. C Credit: Unknown author.{{free media}}
Neodymium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Thorianite contains the oxides of uranium, lanthanum, cerium, praseodymium and neodymium.

"Along with uranium, zinc, iron ore, copper and gold, Greenland’s ancient rocks also harbor large quantities of those minerals known as “rare earth,” among them lanthanum, cerium, neodymium, praesodymium, terbium and yttrium."[51]

Neodymium is in the alloys used to make high-strength neodymium magnets—a type of powerful permanent magnet.[93]

To make neodymium magnets it is alloyed with iron, which is a ferromagnet.[94]

Neodymium magnets, an alloy, Nd2Fe14B, are the strongest permanent magnets known[95] and tend to corrode.[96]

Neons

[edit | edit source]

Neptuniums

[edit | edit source]
File:Neptunium Aeschynite.png
This mineral, Aeschynite, probably contains on the order of a few atoms of neptunium at any one time, as part of the complex decay chain of the uranium that makes up a much larger fraction of the sample. Credit: Theodore Gray.{{fairuse}}
Neptunium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The image at the right shows a rock with the mineral Aeschynite approximately centered above the biotite mica. Aeschynite "probably contains on the order of a few atoms of neptunium at any one time, as part of the complex decay chain of the uranium that makes up a much larger fraction of the sample."[97]

Pure neptunium is paramagnetic, NpAl3 is ferromagnetic, NpGe3 has no magnetic ordering, and NpSn3 behaves fermionically.[98]

Nickels

[edit | edit source]
This piece of Ni is about 3 cm in size. Credit: Materialscientist.{{free media}}
This is an emission-line spectrum for nickel over the visible range: 400-700 nm. Credit: McZusatz.{{free media}}
The image shows the color of nickel in an oxy-hydrogen flame. Credit: NSEasternShoreChemist. {{free media}}

.

Nickel has an emission line occurring in the solar corona at 511.603 nm from Ni XIII.[99]

Nickel has an emission line occurring in the solar corona at 670.183 nm from Ni XV.[99]

Nickel has three emission lines occurring in the solar corona at 380.08 nm of Ni XIII and 423.14 nm and 431.1 of Ni XII.[99]

Nickel has an absorption band at 401.550-436.210 nm with an excitation potential of 4.01 eV.[84]

Antitaenite is a meteoritic metal alloy mineral composed of iron and nickel, 20-40% Ni (and traces of other elements).[100]

Breithauptite is a nickel antimonide mineral with the simple formula NiSb.

Niccolite has the chemical formula NiAs.[19]

Nihoniums

[edit | edit source]

Stability of a nucleus is provided by the strong interaction. However, its range is very short; as nuclei become larger, their influence on the outermost nucleons (protons and neutrons) weakens. At the same time, the nucleus is torn apart by electrostatic repulsion between protons, as it has unlimited range.[101] Nuclei of the heaviest elements are thus theoretically predicted[102] and have so far been observed[103] to primarily decay via decay modes that are caused by such repulsion: alpha decay and spontaneous fission; not all decay modes are caused by electrostatic repulsion. For example, beta decay is caused by the weak interaction.[104] these modes are predominant for nuclei of superheavy elements. Alpha decays are registered by the emitted alpha particles, and the decay products are easy to determine before the actual decay; if such a decay or a series of consecutive decays produces a known nucleus, the original product of a reaction can be determined arithmetically. Since mass of a nucleus is not measured directly but is rather calculated from that of another nucleus, such measurement is called indirect. Direct measurements are also possible, but for the most part they have remained unavailable for heaviest nuclei.[105] The first direct measurement of mass of a superheavy nucleus was reported in 2018 at LBNL.[106] Mass was determined from the location of a nucleus after the transfer (the location helps determine its trajectory, which is linked to the mass-to-charge ratio of the nucleus, since the transfer was done in presence of a magnet).[107] Spontaneous fission, however, produces various nuclei as products, so the original nuclide cannot be determined from its daughters.[108] a leading scientist at JINR, and thus it was a "hobbyhorse" for the facility.[109] In contrast, the LBL scientists believed fission information was not sufficient for a claim of synthesis of an element. They believed spontaneous fission had not been studied enough to use it for identification of a new element, since there was a difficulty of establishing that a compound nucleus had only ejected neutrons and not charged particles like protons or alpha particles.[110] They thus preferred to link new isotopes to the already known ones by successive alpha decays.[108]

Nihonium monofluorides

[edit | edit source]

The analogous monofluoride (NhF) should also exist.[111]

Nihonium monohydrides

[edit | edit source]

The simplest possible nihonium compound is the monohydride, NhH. The bonding is provided by the 7p1/2 electron of nihonium and the 1s electron of hydrogen. The SO interaction causes the binding energy of nihonium monohydride to be reduced by about 1 eV[112] and the nihonium–hydrogen bond length to decrease as the bonding 7p1/2 orbital is relativistically contracted. This is unique among the 7p element monohydrides; all the others have relativistic expansion of the bond length instead of contraction.[113] Another effect of the SO interaction is that the Nh–H bond is expected to have significant pi bonding character (side-on orbital overlap), unlike the almost pure sigma bonding (head-on orbital overlap) in thallium monohydride (TlH).[114]

Nihonium monoiodides

[edit | edit source]

Nihonium(I) is predicted to be more similar to silver(I) than thallium(I):[112] the Nh+ ion is expected to more willingly bind anions, so that NhCl should be quite soluble in excess hydrochloric acid or ammonia; thallium(I) chloride (TlCl) is not. In contrast to Tl+, which forms the strongly basic hydroxide (thallium(I) hydroxide (TlOH)) in solution, the Nh+ cation should instead hydrolyse all the way to the amphoteric oxide Nh2O, which would be soluble in aqueous ammonia and weakly soluble in water.[115]

Niobiums

[edit | edit source]
File:Iron-niobium phase diagram.png
This is an iron-niobium phase diagram. Credit: E. Paul and L.J. Swartendruber.{{fairuse}}
High purity (99.995 % = 4N5) niobium crystals, electrolytic made, as well as a high purity (99.95 % = 3N5) 1 cm3 anodized niobium cube for comparison. Credit: Alchemist-hp.{{free media}}
Niobium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

As can be seen in the iron-niobium phase diagram on the left, niobium is single phase (α-Nb) up to its melting temperature. This is a bcc structure.

Ferroniobium is an alloy of 60–70% niobium with iron, where niobium is used mostly in alloy steel.[116][117][118] Niobium is used in various superconducting materials, Type-II superconductor alloys, also containing titanium and tin. Quantities of niobium are used in nickel-, cobalt-, and iron-based superalloys in proportions as great as 6.5%.[119]

It appears to be the case that native niobium does not occur in the surface rocks on Earth.

Nitrogens

[edit | edit source]

Carlsbergites

[edit | edit source]
Agpalilik meteorite is outside the Geological Museum in Copenhagen. Credit: Michael B. H..{{free media}}

Carlsbergite was first described in the Agpalilik fragment of the Cape York meteorite.

It is a chromium nitride mineral (CrN),[120] named after the Carlsberg Foundation that backed the recovery of the Agpalilik fragment from the Cape York meteorite.[120]

It occurs in meteorites along the grain boundaries of kamacite or troilite in the form of tiny plates,[120] associated with kamacite, taenite, daubreelite, troilite and sphalerite.[121]

Nobeliums

[edit | edit source]

A nobelium atom has 102 electrons, of which three can act as valence electrons. They are expected to be arranged in the configuration [Rn]5f147s2 (ground state term symbol 1S0), although experimental verification of this electron configuration had not yet been made as of 2006.[122] In forming compounds, all the three valence electrons may be lost, leaving behind a [Rn]5f13 core: this conforms to the trend set by the other actinides with their [Rn]5fn electron configurations in the tripositive state. Nevertheless, it is more likely that only two valence electrons may be lost, leaving behind a stable [Rn]5f14 core with a filled 5f14 shell. The first ionization potential of nobelium was measured to be at most (6.65 ± 0.07) eV in 1974, based on the assumption that the 7s electrons would ionize before the 5f ones;[123] this value has not yet been refined further due to nobelium's scarcity and high radioactivity.[124] The ionic radius of hexacoordinate and octacoordinate No3+ had been preliminarily estimated in 1978 to be around 90 and 102 pm respectively;[122] the ionic radius of No2+ has been experimentally found to be 100 pm to two significant figures.[122] The enthalpy of hydration of No2+ has been calculated as 1486 kJ/mol.[122]

Oganessons

[edit | edit source]

Although oganesson is a member of group 18 (the noble gases) – the first synthetic element to be so – it may be significantly reactive, unlike all the other elements of that group.[125] It was formerly thought to be a gas under standard conditions for temperature and pressure but is now predicted to be a solid due to relativistic effects.[125]

Osmiums

[edit | edit source]
File:Native Osmium crystal.png
A crystal of native osmium from Nishni Tagil, Ural, Russia, is shown. Credit: Thomas Witzke.{{fairuse}}
Osmium Os shown are crystals, purity ≥ 99.99%, 2.2 g. Credit: Alchemist-hp.{{free media}}
Osmium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The crystal of native osmium shown on the right is about 2 mm across.

Osmium alloys with platinum, iridium, and other platinum-group metals.[126]

Oxygens

[edit | edit source]

Palladiums

[edit | edit source]
File:Palladium nugget Brazil.png
This is a palladium nugget. Credit: Hudson Institute of Mineralogy. {{fairuse}}
File:Palladium Mednorudyanskoye Cu Deposit.jpg
This piece of native palladium is from the Mednorudyanskoye Cu Deposit, Nizhnii Tagil, Sverdlovskaya Oblast', Middle Urals, Urals Region, Russia. Credit: Hudson Institute of Mineralogy.{{fairuse}}
Palladium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

"Natural Palladium [like the nugget shown on the right] always contains some Platinum."[127]

This palladium nugget is from Bom Sucesso Creek, Serro, Minas Gerais, Brazil.

"(Pd,Cu) alloys, some with the approximate composition PdCu4, are reported by Kapsiotis et al. (2010)."[127]

The piece of native palladium [image on the left] from the Mednorudyanskoye Cu Deposit, Nizhnii Tagil, Sverdlovskaya Oblast', Middle Urals, Urals Region, Russia, probably contains some copper.

Palladium can be found as a free metal alloyed with gold and other platinum-group metals in placer deposits of the Ural Mountains, Australia, Ethiopia, North and South America.

Palladium is found in the rare minerals cooperite[128] and polarite.[129] Many more Pd minerals are known, but all of them are very rare.[130]

Potarites

[edit | edit source]
File:Potarite from Brazil.jpg
This piece of potarite is from Serro, Minas Gerais, Brazil. Credit: Hudson Institute of Mineralogy.{{fairuse}}

Potarite has the chemical formula PdHg.[19]

On the right is a piece of potarite is from Serro, Minas Gerais, Brazil.

Phosphoruses

[edit | edit source]
White phosphorus and resulting allotropes, including violet phosphorus, are indicated. Credit: UserXresu.{{free media}}
White phosphorus is under water on the left, with red phosphorus (center images), and violet phosporus right. Credit: Materialscientist.{{free media}}
Phosphorus spectrum is for emission lines between 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Phosphorus has several allotropes that exhibit strikingly diverse properties.[131] The two most common allotropes are white phosphorus and red phosphorus.[132]

Violet phosphorus is a form of phosphorus that can be produced by day-long annealing of red phosphorus above 550 °C, when phosphorus was recrystallised from molten lead, a red/purple form is obtained, sometimes known as "Hittorf's phosphorus" (or violet or α-metallic phosphorus).[133]

"It would appear that violet phosphorus is a polymer of high relative molecular mass, which on heating breaks down into P2 molecules. On cooling, these would normally dimerize to give P4 molecules (i.e. white phosphorus) but, in vacuo, they link up again to form the polymeric violet allotrope."[134]

Phosphorus is an important component in steel production, in the making of phosphor bronze, and in many other related products.[135][136]

Phosphorus is added to metallic copper during its smelting process to react with oxygen present as an impurity in copper and to produce phosphorus-containing copper (CuOFP) alloys with a higher hydrogen embrittlement resistance than normal copper.[137]

Platinums

[edit | edit source]
This is a native platinum nugget, locality Kondyor mine, Khabarovsk Krai, Russia. Credit: Alchemist-hp.{{fairuse}}
Crystals of pure platinum were grown by gas phase transport. Credit: Periodictableru.{{free media}}
Platinum spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Platinum can also occur as nuggets such as the one imaged on the right from Russia.

"Terrestrial iron-free rhodium-bearing platinum with the composition of Pt0.68Rh0.32 in association with platinum-bearing rhodium Rh0.57Pt0.43 [...] was originally discovered in heavy fractions from basic rocks (norite, gabbro, and anorthosite) in the upper zone of the layered Stillwater intrusion (Montana, United States) [2]."[138]

In nickel and copper deposits, platinum-group metals occur as sulfides (e.g., (Pt,Pd)S), tellurides (e.g., PtBiTe), antimonides (PdSb), and arsenides (e.g. PtAs2), and as end alloys with nickel or copper.

Cooperites

[edit | edit source]
Outstanding specimen is the extremely rare platinum mineral cooperite (PtS). Credit: David Hospital.{{free media}}

This specimen on the right is a single nugget from Tulameen River, Princeton, British Columbia, Canada.

Plutoniums

[edit | edit source]
File:Muromontite.png
This is a sample of the naturally occurring mineral muromontite. Credit: eBay seller rubbleshop.{{fairuse}}
Plutonium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

"The sample [on the right] representing plutonium is the naturally occurring mineral muromontite, which is a mixture of uranium and beryllium. [The] alpha particles from the decay of uranium are captured by the beryllium atoms, which in turn release neutrons. [...] In the case of this sample, [...] the neutrons are in turn re-captured by the uranium, which then undergoes further decay and is transformed into plutonium. The result is that this mineral contains the highest known naturally occurring concentration of plutonium."[139]

Gallium, aluminium, americium, scandium and cerium can stabilize the δ phase of plutonium for room temperature, silicon, indium, zinc and zirconium allow formation of metastable δ state when rapidly cooled, high amounts of hafnium, holmium and thallium also allows some retention of the δ phase at room temperature, but nNeptunium is the only element that can stabilize the α phase at higher temperatures.[140]

Poloniums

[edit | edit source]
File:Polonium Halo in Biotite.png
This photograph shows a 210Po halo in biotite from the Buckhorn pegmatite. Credit: Lorence G. Collins.{{fairuse}}
Uranium roll front occurs in quartzose sandstone in the Cretaceous of Colorado, USA. Credit: James St. John.{{free media}}
File:Radioactive decay halos along crack.png
This photo shows a fracture in biotite in which migrating 210Po and/or 210Pb ions have created damage to the biotite lattice parallel to the fracture. Credit: Lorence G. Collins.{{fairuse}}
Polonium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

α-Po crystallizes in a simple cubic lattice.[141]

Native polonium may occur in minerals like pitchblende due to the decay of uranium. But, when the uranium is chemically bound, the polonium is likely to be also.

β-Po has a rhombohedral (trigonal) crystal structure.[142]

"Solid diorite and gabbro rock, which had previously crystallized from magma, has been subjected to repeated cataclasis and recrystallization. This has happened without melting; and the cataclasis provided openings for the introduction of uranium-bearing fluids and for the modification of these rocks to granite by silication and cation deletion."[143]

"In uranium ore-fields the extra uranium provides an abundant source of inert radon gas; and it is this gas that diffuses in ambient fluids so that incipient biotite and fluorite crystallization is exposed to it. Radon (222Rn) decays and Po isotopes nucleate in the rapidly growing biotite (and fluorite) crystals whence they are positioned to produce the Po halos."[143]

On the lower right is a photograph showing radioactive decay halos along a crack in biotite.

On the left is an example of groundwater incursion that has moved through a nearby fault. The groundwater has picked up dissolved uranium compounds and moved downward through adjacent porous sandstones. Uraninite then precipitated around a tongue of groundwater, resulting in the roll front seen in the image on the left.

Potassiums

[edit | edit source]
This is a pressure-temperature phase diagram for potassium. Credit: David A. Young, ERDA.{{fairuse}}
Cut potassium (kalium) pieces show metallic silver color. Credit: Dennis s.k.{{free media}}
File:Potassium flame.png
The image shows the color of potassium in a natural gas burner. Credit: SciencePhoto. {{fairuse}}
This is a potassium emission-line spectrum over the range 400 nm - 700 nm. Credit: McZusatz.{{free media}}

As indicated in the phase diagram on the left, potassium occurs in a bcc (α-K) phase from room temperature up to melting.

Native potassium does not appear to occur on the Earth's surface.

"The group [of potassium lines] at λλ 535, 510, and 495 Å showed no trace of structure even in an arc of but half an ampere."[144]

Praseodymiums

[edit | edit source]
Praseodymium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Po-BeO mixtures or alloys used as neutron sources are a neutron trigger or initiator for nuclear weapons[53][145] and for inspections of oil wells.

The polonide of praseodymium (PrPo) melts at 1250 °C, and that of thulium (TmPo) melts at 2200 °C.[39] PbPo is one of the very few naturally occurring polonium compounds, as polonium alpha decays to form lead.[146]

Promethiums

[edit | edit source]
Promethium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Formula symmetry space group No Pearson symbol a (pm) b (pm) c (pm) Z density,
g/cm3
α-Pm Close-packing of equal spheres (dhcp)[147][148] P63/mmc 194 hP4 365 365 1165 4 7.26
β-Pm Cubic crystal system (bcc)[148] Fm3m 225 cF4 410 410 410 4 6.99

Protactiniums

[edit | edit source]
File:Protactinium.jpg
Protactinium crystal, prepared by the van Arkel (chemical vapour transport) process. Credit: The Actinide Group, the Institute for Transuranium Elements.{{fairuse}}
This sample of Protactinium-233 (dark circular area in the photo) was photographed in the light from its own radioactive emission (the lighter area) at the National Reactor Testing Station in Idaho. Credit: ENERGY.GOV.{{free media}}
Protactinium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Protactinium is one of the rarest and most expensive naturally occurring elements, found in the form of two isotopes – 231Pa and 234Pa, with the isotope 234Pa occurring in two different energy states: nearly all natural protactinium is protactinium-231, an alpha emitter formed by the decay of uranium-235, whereas the beta radiating protactinium-234 is produced as a result of uranium-238 decay. Nearly all uranium-238 (99.8%) decays first to the shorter-lived 234mPa isomer.[149]

Radiums

[edit | edit source]
Radium electroplated on a very small sample of copper foil and covered with polyurethane to prevent reaction with the air. Credit: grenadier.{{free media}}
Radium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
File:Radium flame.png
The image shows the color of radium in a natural gas burner. Credit: Alan Crooks. {{fairuse}}

"Solid radium is bcc at room temperature. Radium melts at 973 K.63"[150]

Radium oxide (RaO) has not been characterized well past its existence, despite oxides being common compounds for the other alkaline earth metals.

Radons

[edit | edit source]

Rheniums

[edit | edit source]
File:Native rhenium 2.png
This is an image of the first find of native rhenium in the transitional clay layer at the Cretaceous/Paleogene boundary in the Gams Section. Credit: A. F. Grachev, S. E. Borisovsky, and A. V. Grigor’eva.{{fairuse}}
Three high-purity forms of rhenium metal are shown: a single crystal (99.999% pure) made by the floating-zone process, an e-beam remelted bar (99.995% pure), and a 1 cm3 cube (99.99% pure) for comparison. Credit: Alchemist-hp.{{free media}}
Rhenium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

"Native rhenium was first discovered in the Earth's crust in wolframites from a rare metal deposit in the Transbaikal region [1]. [...] The study of the lunar regolith from two sites revealed native rhenium particles with different morphological features: irregular dense particles from Mare Fecunditatis and spheroidal particles from Mare Crisium. The origin of particles (less than 10 µm in size) is assigned to exhalative processes [2]. Among the extraterrestrial objects, native rhenium was found in Ni-iron and silicates from the Allende meteorite [3]."[151]


"Solid radium is bcc at room temperature. Radium melts at 973 K.63"[150]

"Instrumental neutron activation analyses of Kilauean aerosols collected in 1984 show Ir:Au:Re ratios of 1:12:2000 normalized to CI chondrites. The large Re enrichment in these volcanic aerosols may explain the 3 to 15-fold Re excess, relative to chondritic, in the observed siderophile element signature of the Cretaceous-Tertiary boundary clay layer.2,3 Strong evidence exists that an impact of an extraterrestrial body on the Earth caused mass extinctions at the end of the Cretaceous period.4,5 ... A Kilauean aerosol contribution of only 0.01 % of the chondritic component in the boundary clay layer would produce the observed Re enrichments."[152]

Rhodiums

[edit | edit source]
File:Native rhodium.jpg
This image of microbreccia contains rhodium-bearing ferroplatinum. Credit: T. A. Gornostaeva, P. M. Kartashov, A. V. Mokhov, and O. A. Bogatikov.{{fairuse}}
Rhodium processing is shown: 1g powder, 1g pressed cylinder, 1 g argon arc remelted pellet. Credit: Alchemist-hp.{{free media}}
Rhodium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The image on the right contains small particles of native rhodium-bearing ferroplatinum. This sample was obtained from the lunar regolith "by the Luna-16 automatic station".[138]

"Terrestrial iron-free rhodium-bearing platinum with the composition of Pt0.68Rh0.32 in association with platinum-bearing rhodium Rh0.57Pt0.43 [...] was originally discovered in heavy fractions from basic rocks (norite, gabbro, and anorthosite) in the upper zone of the layered Stillwater intrusion (Montana, United States) [2]."[138]

Naturally occurring rhodium is usually found as a free metal or as an alloy with similar metals and rarely as a chemical compound in minerals such as bowieite and rhodplumsite.

Rhodium is used as an alloying agent for hardening and improving the corrosion resistance.[153] of platinum and palladium.

Roentgeniums

[edit | edit source]

"Based on the observation of the long-lived isotopes of roentgenium, 261Rg and 265Rg (Z = 111, t1/2 ≥ 108 y) in natural Au, an experiment was performed to enrich Rg in 99.999% Au. 16 mg of Au were heated in vacuum for two weeks at a temperature of 1127°C (63°C above the melting point of Au). The content of 197Au and 261Rg in the residue was studied with high resolution inductively coupled plasma-sector field mass spectrometry (ICP-SFMS). The residue of Au was 3 × 10−6 of its original quantity. The recovery of Rg was a few percent. The abundance of Rg compared to Au in the enriched solution was about 2 × 10−6, which is a three to four orders of magnitude enrichment."[154]

The isotopes 280Rg and 281Rg are promising for chemical experimentation and may be produced as the granddaughters of the moscovium isotopes 288Mc and 289Mc respectively;[47] their parents are the nihonium isotopes 284Nh and 285Nh, which have already received preliminary chemical investigations.[155]

Rubidiums

[edit | edit source]
This is a pressure-temperature phase diagram for rubidium. Credit: David A. Young, ERDA.{{free media}}
Rubidium metal sample is enclosed in a quartz ampule to prevent oxidation. Credit: .{{free media}}
This image shows the flame test for Rubidium. Credit: Didaktische.Medien.{{free media}}
Rubidium emission spectrum is for 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The pressure-temperature diagram on the left shows that rubidium is bcc (α-Rb) from room temperature through melting.

Native rubidium does not appear to occur on the Earth's surface.

Rutheniums

[edit | edit source]
File:Native Ruthenium from Verkneivinsk.png
This is a piece of native ruthenium from Verkhneivinsk, Neiva river, Sverdlovskaya Oblast', Middle Urals, Urals Region, Russia. Credit: Hudson Institute of Mineralogy.{{fairuse}}
One half of a high-purity (99.99%), electron-beam-remelted ruthenium bar is shown. Credit: Alchemist-hp.{{free media}}
Gas phase grown crystals of ruthenium metal are shown. Credit: Periodictableru.{{free media}}
Ruthenium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The piece of native ruthenium in the image on the right contains some iridium. It is from Verkhneivinsk, Neiva river, Sverdlovskaya Oblast', Middle Urals, Urals Region, Russia.

A minor application for ruthenium is in platinum alloys. A ruthenium-molybdenum alloy is known to be superconductive at temperatures below 10.6 K.[156] The composition of the mined platinum group metal (PGM) mixtures varies widely, depending on the geochemical formation. For example, the PGMs mined in South Africa contain on average 11% ruthenium while the PGMs mined in the former USSR contain only 2% (1992).[157][158]

Rutherfordiums

[edit | edit source]

Rutherfordium was synthesized by bombarding a californium-249 target with carbon-12 ions and measured the alpha decay of 257Rf, correlated with the daughter decay of nobelium-253:[159]

249
98
Cf
+ 12
6
C
257
104
Rf
+ 4 n

Rutherfordium is the parent of K-alpha X-rays in the elemental signature of the 257Rf decay product, nobelium-253.[160]

Samariums

[edit | edit source]
Ultrapure sublimated samarium, 2 grams, is shown. Credit: Unknown author.{{free media}}
Samarium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Samarium occurs in concentration up to 2.8% in several minerals including cerite, gadolinite, samarskite, monazite and bastnäsite.

Scandiums

[edit | edit source]
Scandium emission spectrum is from 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Scandium is the first transition metal and the first rare earth element. Credit: Hi-Res Images of Chemical Elements.{{free media}}

Scandium is the first transition metal and the first rare earth element, the latter also includes yttrium and the lanthanoids. The ignoble light metal has only a few applications, because its chemistry isn't so complex and it also is rather expensive. It is used in high-quality, light alloys, e.g., for frames of racing bicycles.

Scandium (Sc II) has an absorption band, 424.683±1.0 nm, with an excitation potential of 0.31 eV.[84]

Metallic scandium is used in aluminium alloysis for strengthening with as little as 0.5% scandium.[167][168]

The alloy Al
20
Li
20
Mg
10
Sc
20
Ti
30
is as strong as titanium, light as aluminium, and hard as some ceramics.[169]

"Neutron activation analysis was used to deterimne the total [lanthanum] La and [scandium] Sc content of three soils developed from loess-capped glacial till. The profiles were classified as Gray-Brown Podzolics (Hapludalfs) overlying paleosols developed in Rockain till. The total La content in the less than 250µ fraction of these soils ranged from 18.1 to 37.1 ppm, with an average content of 23.7 ppm in the loess and 28.5 ppm in the glacial till. Total Sc in the soils ranged from 5.1 to 10.9 ppm with average contents of 6.5 and 9.0 ppm in the loess and glacial till, respectively. Translocation by pedogenic processes was indicated by the accumulation of these elements in the argillic B horizons. Correlation coefficients of La and Sc with clay percentages in the profiles were 0.79 and 0.88, respectively."[170]

Seaborgiums

[edit | edit source]

Alpha emission (α), spontaneous fission (SF) and electron capture (EC) are decay modes of seaborgium.

List of seaborgium isotopes
Isotope
Half-life
[171][172]
Decay
mode[171][172]
Discovery
year
Reaction
258Sg 3 ms SF 1994 209Bi(51V,2n)
259Sg 600 ms α 1985 207Pb(54Cr,2n)
260Sg 4 ms SF, α 1985 208Pb(54Cr,2n)
261Sg 200 ms α, EC, SF 1985 208Pb(54Cr,n)
261mSg 92 μs IT 2009 208Pb(54Cr,n)
262Sg 7 ms SF, α 2001 270Ds(—,2α)
263Sg 1 s α 1994 271Ds(—,2α)
263mSg 120 ms α, SF 1974 249Cf(18O,4n)
264Sg 37 ms SF 2006 238U(34Si,4n)
265Sg 8 s α 1993 248Cm(22Ne,5n)
265mSg 16.2 s α 1993 248Cm(22Ne,5n)
266Sg 360 ms SF 2004 270Hs(—,α)
267Sg 1.4 min SF, α 2004 271Hs(—,α)
269Sg 14 min α 2010 285Fl(—,4α)
271Sg 2.4 min α 2003 287Fl(—,4α)

Seleniums

[edit | edit source]
Selenium (native) with pen for scale is from the mineral collection of Brigham Young University Department of Geology, Provo, Utah. Credit: Andrew Silver, USGS.{{free media}}
The dark gray mineral in the yellow sandstone is native selenium. Credit: James St. John.{{free media}}
File:Native selenium.jpg
These are native selenium needles from Katharine mine, Radvanice, Czech Republic. Credit: Asahi.{{fairuse}}
Black, glassy amorphous (with thin layer of grey selenium) and red amorphous selenium are allotropes of selenium. Credit: W. Oelen.{{free media}}
Selenium emission spectrum for 400 nm - 700 nm. Credit: McZusatz.{{free media}}

On the right is a photograph of native selenium from the mineral collection of Brigham Young University Department of Geology, Provo, Utah.

The image on the left shows dark gray selenium in sandstone from Westwater Canyon Section 23 Mine Grants, New Mexico.

In the center image are native selenium needles from Katharine mine, Radvanice, Czech Republic.

Allotropes of selenium are amorphous, brick-red (α, β,[173][174] and γ[175]) powders, black, vitreous beads,[176] and gray selenium.

Selenium is used with bismuth in brasses to replace lead.[177] Like lead and sulfur, selenium improves the machinability of steel at concentrations around 0.15%.[178][179] Selenium produces the same machinability improvement in copper alloys.[177]

Achávalites

[edit | edit source]
Black metallic crystals are the extremely rare iron selenide mineral achávalite from Cacheuta Mine, Mendoza, Argentina, associated with cacheutaite. Credit: David Hospital.{{free media}}

Achávalite has the chemical formula (Fe,Cu)Se.

Achávalite (IMA symbol is Ahv[180]) a selenide mineral that is a member of the nickeline group. It has only been found in a single Argentinian mine system, being first discovered in 1939 in a selenide deposit. The type locality is the Cacheuta mine, Sierra de Cacheuta, Mendoza, Argentina.[181][182][183]

Clausthalites

[edit | edit source]
Clausthalite is a rare lead selenide exhibiting scintillating, metallic microcrystals covering the carbonate matrix. Credit: Rob Lavinsky.{{free media}}

Clausthalite is a lead selenide mineral, with chemical formula PbSe. It is a face-centered mineral with Z = 4 formula units per unit cell.

Siderophiles

[edit | edit source]
Abundance (atom fraction) of the chemical elements in Earth's upper continental crust as a function of atomic number. The rarest elements in the crust (shown in yellow) are not the heaviest, but are rather the siderophile (iron-loving) elements in the Goldschmidt classification of elements. These have been depleted by being relocated deeper into the Earth's core. Their abundance in meteoroid materials is relatively higher. Additionally, tellurium and selenium have been depleted from the crust due to formation of volatile hydrides. Credit: Gordon B. Haxel, Sara Boore, and Susan Mayfield from USGS.{{free media}}

Def. "an element that forms alloys easily with iron and [may be] concentrated in the Earth's core"[184] is called a siderophile.

Siderophile (metal-loving) chemical elements include W, P, Co, Ni, Mo, Re, and Ir.[185]

"The platinum group elements (PGE: Os, Ir, Ru, Rh, Pt, and Pd) and Re are highly siderophile elements (HSE)".[186]

"We believe that silicon is a major element - about 5% [of the Earth's inner core] by weight could be silicon dissolved into the iron-nickel alloys."[187]

"The innermost part of Earth is thought to be a solid ball with a radius of about 1,200 km (745 miles)."[188]

"It is mainly composed of iron, which makes up an estimated 85% of its weight, and nickel, which accounts for about 10% of the core."[188]

"These difficult experiments are really exciting because they can provide a window into what Earth's interior was like soon after it first formed, 4.5 billion years ago, when the core first started to separate from the rocky parts of Earth."[189]

"But other workers have recently suggested that oxygen might also be important in the core."[189]

"In a way, these two options [oxygen was sucked into the core that would leave the rocky mantle surrounding the core depleted of the element or a larger amount of silicon had been incorporated in Earth's core more than four billion years ago, that would have left the rest of the planet relatively oxygen rich] are real alternatives that depend a lot on the conditions prevailing when Earth's core first began to form."[189]

Silicons

[edit | edit source]
Close up photo is of a piece of purified silicon. Credit: Enricoros.{{free media}}
Silicon spectrum is from 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Silicon (Si II) has two absorption bands at 412.805±10.8 nm and 413.088±13.0 nm with excitation potentials of 9.79 eV and 9.80 eV, respectively.[84]

Silicon has an absorption line (Si IV) at 408.9 nm.[190]

Elemental silicon is added to molten cast iron as ferrosilicon or silicocalcium alloys to improve performance in casting thin sections and to prevent the formation of cementite where exposed to outside air. The presence of elemental silicon in molten iron acts as a sink for oxygen, so that the steel carbon content, which must be kept within narrow limits for each type of steel, can be more closely controlled. Ferrosilicon production and use is a monitor of the steel industry, and although this form of elemental silicon is grossly impure, it accounts for 80% of the world's use of free silicon. Silicon is an important constituent of electrical steel, modifying its resistivity and ferromagnetic properties.

The properties of silicon may be used to modify alloys with metals other than iron. "Metallurgical grade" silicon is silicon of 95–99% purity. About 55% of the world consumption of metallurgical purity silicon goes for production of aluminium-silicon alloys (silumin alloys) for aluminium part casts, mainly for use in the automotive industry. Silicon's importance in aluminium casting is that a significantly high amount (12%) of silicon in aluminium forms a eutectic mixture which solidifies with very little thermal contraction. This greatly reduces tearing and cracks formed from stress as casting alloys cool to solidity. Silicon also significantly improves the hardness and thus wear-resistance of aluminium.[191][192]

"The relatively long-lived radionuclide of silicon, 32Si, finds important applications as a tracer for studying aqueous geochemistry, biogeochemical cycles of silicon in the oceans, and the chronology of glaciers and biogenic silica-rich sediments in lacustrine and marine environments."[193]

Silicon hydrogenated amorphous carbons

[edit | edit source]

"The broad, 60 < FWHM < 100 nm, featureless luminescence band known as extended red emission (ERE) is seen in such diverse dusty astrophysical environments as reflection nebulae17, planetary nebulae3, HII regions (Orion)12, a Nova11, Galactic cirrus14, a dark nebula7, Galaxies8,6 and the diffuse interstellar medium (ISM)4. The band is confined between 540-950 nm, but the wavelength of peak emission varies from environment to environment, even within a given object. ... the wavelength of peak emission is longer and the efficiency of the luminescence is lower, the harder and denser the illuminating radiation field is13. These general characteristics of ERE constrain the photoluminescence (PL) band and efficiency for laboratory analysis of dust analog materials."[194]

"The PL efficiencies measured for [hydrogenated amorphous carbon] HAC and Si-HAC alloys are consistent with dust estimates for reflection nebulae and planetary nebulae, but exhibit substantial photoluminescence below 540 nm which is not observed in astrophysical environments."[194]

Moissanite

[edit | edit source]
Moissanite is native SiC. Credit: Andrew Silver.{{free media}}

Moissanite is native SiC.[19]

Silvers

[edit | edit source]
These specimens are some of the most easily recognizable, dramatic and highly sought after silver specimens from the Western Hemisphere. Credit: Robert Lavinsky.{{free media}}
Native silver is on calcite from Kongsberg, Norway. Credit: Aram Dulyan.{{fairuse}}
A pure (>99.95%) silver crystal, synthetic electrolytic is made with visible dendritic structures. Credit: Alchemist-hp.{{free media}}
Silver spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Native silver does occur as cubic, octahedral, or dodecahedral crystals; "also elongated, arborescent, reticulated, or as thin to thick wires."[19]

The metal is found in the Earth's crust in the pure, free elemental form ("native silver"), as an alloy with gold, copper, zinc, cadmium, indium, tin, mercury, cobalt, nickel, palladium, manganese, phosphorus, and lead, and in minerals such as argentite and chlorargyrite.

Ag+ is the stable species in aqueous solution and solids.[39]

Bromyrites

[edit | edit source]
File:Bromyrite.jpg
This is a butterscotch colored bromargyrite cube from Broken Hill, New South Wales, Australia. Credit: Lou Perloff / Photo Atlas of Minerals.{{fairuse}}

Bromyrite, or bromargyrite, is a cubic silver bromide mineral (AgBr) that is 50 at % bromine.

The image on the right shows a butterscotch colored bromargyrite cube from Broken Hill, New South Wales, Australia.

Iodyrites

[edit | edit source]
File:Iodyrite.jpg
These are twinned iodyrite, or iodargyrite, crystals. Credit: Hudson Institute of Mineralogy.{{fairuse}}

Iodyrite (AgI) may be the most common mineral with large amounts of iodine found on Earth. It is 50 at % iodine.

On the right are twinned iodyrite, or iodargyrite, crystals are within a rock sample from Schöne Aussicht Mine, Dernbach, Neuwied, Wied Iron Spar District, Westerwald, Rhineland-Palatinate, Germany.

Sodiums

[edit | edit source]
Sodium spectrum is ringrising 400 nm - 700 nm Credit: McZusatz.{{free media}}
File:Bi-Na phase diagram.gif
This is a phase diagram for the sodium-bismuth system. Credit: J. Sangster and A.D. Pelton.{{fairuse}}
Sodium metal is from the Dennis s.k collection. Credit: Dennis s.k.{{free media}}
The image shows the color of sodium in a natural gas burner. Credit: Søren Wedel Nielsen.{{free media}}

.

The phase diagram on the left shows bcc (α-Na) at higher temperatures up to melting and hcp (β-Na) with decreasing temperature below the transition at 97.8°C.

Native sodium does not appear to occur on the surface of the Earth.

"Glaciers in the Karakoram and western Himalaya (site 2 and 3) show high annual snow accumulation rates and high annual fluxes of calcium, sodium, chloride, sulfate, and nitrate."[195]

Fluorites

[edit | edit source]
These are cyan colored fluorite crystals from Rogerley Mine, Frosterley, Weardale, North Pennines, Co. Durham, England, UK. Credit: Parent Géry.{{free media}}
The fluorite crystal is just over 1 cm and is rimmed on one side with sparkling pyrite. Credit: Robert Lavinsky.{{free media}}

Fluorite is a mineral composed of NaF.

Although fluorite usually appears violet or purple in color, the crystals at left are cyan with some blue or violet fluorite mixed in suggesting slight variations in composition.

Strontiums

[edit | edit source]
File:Magnesium-Strontium phase diagram.png
This is a magnesium-strontium phase diagram. Credit: A.A. Nayeb-Hashemi and J.B. Clark.{{fairuse}}
The image shows the color of strontium in a natural gas burner. Credit: Herge.{{free media}}
The image shows the color of strontium in a natural gas burner. Credit: Roman Kynčl.{{free media}}
The chemical element strontium as synthetic crystals, sealed under argon in a glass ampoule, purity (99.95 %). Credit: Alchemist-hp.{{free media}}
Strontium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Strontium at room temperature crystallizes in a fcc structure (α-Sr).

According to the phase diagram on the left, α-Sr transforms to γ-Sr (bcc) at 547°C.

Native strontium does not appear to occur on the surface of the Earth.

Three allotropes of metallic strontium exist, with transition points at 235 and 540 °C.[196]

Strontium (Sr II) has two absorption bands: 407.771±11.3 nm and 421.552±10.4 nm.[84]

Sulfurs

[edit | edit source]

Tantalums

[edit | edit source]
File:Native tantalum from Greenland.jpg
This is a piece of native tantalum from Kvanefjeld Mountain, Kuannersuit Plateau, Ilímaussaq complex, Narsaq, Kujalleq, Greenland. Credit: V.V. Seredin.{{fairuse}}
File:Iron-Tantalum phase diagram.png
This is a National Bureau of Standards phase diagram for Fe-Ta. Credit: L.J. Swartzendruber and E. Paul.{{fairuse}}

The iron-tantalum phase diagram on the left shows the bcc (α-Ta) phase from lower temperatures through and up to melting.

On the right is a scanning electron micrograph of a piece of native tantalum from Kvanefjeld Mountain, Kuannersuit Plateau, Ilímaussaq complex, Narsaq, Kujalleq, Greenland.

A high purity (99.999 %) tantalum single crystal, made by the floating zone process, some single crystalline fragments of tantalum, as well as a high purity (99.99 % = 4N) 1 cm3 tantalum cube for comparison. Credit: Alchemist-hp.{{free media}}
Tantalum spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Tantalum forms compounds in oxidation states −III to +V.

A tantalum-tellurium alloy forms quasicrystals.[197]

Technetiums

[edit | edit source]
Technetium sample inside a sealed glass ampoule, filled with argon gas. 6x1 mm goldfoil covered with 99Tc powder (electroplated). Credit: Marco Cardin.{{free media}}
Technetium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
and

These "reactions probe precisely the time scale and neutrino-flux component of most interest: the boron-8 neutrino luminosity, which is the most sensitive monitor of variations in the solar core temperature, during and before the Pleistocene epoch. (The half-lives of technetium-97 and -98 are, respectively, 2.6 and 4.2 million years; the reaction on molybdenum-98 is induced only by the high-energy boron-8 neutrinos; and the reaction on molybdenum-97 may sample in addition the flux of beryllium-7 neutrinos, which are second only to boron-8 neutrinos in sensitivity to the core temperature.)"[198]

Telluriums

[edit | edit source]
This is a native tellurium crystal from the Emperor Mine, Vatukoula, Tavua Gold Field, Viti Levu, Fiji. Credit: Robert Stravinsky. {{free media}}
File:Native tellurium.JPG
On the upper left of the rock is native tellurium. Credit: Theodore W. Gray.{{fairuse}}

On the right is an example of native tellurium from the Emperor Mine, Vatukoula, Tavua Gold Field, Viti Levu, Fiji.

On the left is an encrustation of native tellurium on the upper left portion of a rock.

Metallic tellurium is diameter 3.5 cm. Credit: Unknown author.{{free media}}

Tellurium is used in iron, stainless steel, copper, lead alloys, n-type bismuth telluride alloys[199].

Native tellurium crystal on sylvanite (Vatukoula, Viti Levu, Fiji), picture width 2 mm. Credit: Christian Rewitzer.{{free media}}
Tellurium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Tellurium has two allotropes, crystalline and amorphous. When crystalline, tellurium is silvery-white with a metallic luster. The crystals are trigonal and chiral (space group 152 or 154 depending on the chirality), like the gray form of selenium. It is a brittle and easily pulverized metalloid. Amorphous tellurium is a black-brown powder prepared by precipitating it from a solution of tellurous acid or telluric acid (Te(OH)6).[200]

Altaites

[edit | edit source]
Rich silvery veins and flecks are altaite throughout the matrix. Credit: Rob Lavinsky.{{free media}}

Altaite has the chemical formula of PbTe. It has face-centered cubic structure with four formula molecules (Z=4) per unit cell. It is 50 atomic percent lead and 50 at. % tellurium. Crystal habits include cubic and octahedral crystals; but much more commonly found in massive and granular forms.

Tennessines

[edit | edit source]
Decay chain of the atoms produced in the original experiment to produce tennessine. Credit: See authors.{{free media}}

The figures near the arrows describe experimental (black) and theoretical (blue) values for the lifetime and energy of each decay. Lifetimes may be converted to half-lives by multiplying by ln 2.[201]

Terbiums

[edit | edit source]
Terbium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Terbium is a relatively resistant, sparsely toxic lanthanoid. Credit: Hi-Res Images of Chemical Elements.{{free media}}

Terbium is never found in nature as a free element, but it is contained in many minerals, including cerite, gadolinite, monazite, xenotime and euxenite.

Terfenol-D, an alloy of the formula Tb
x
Dy
1-x
Fe
2
(x ≈ 0.3), is a magnetostrictive material.

Terbium is contained along with other rare earth elements in many minerals, including monazite ((Ce,La,Th,Nd,Y)PO
4
with up to 0.03% terbium), xenotime (YPO
4
) and euxenite ((Y,Ca,Er,La,Ce,U,Th)(Nb,Ta,Ti)
2
O
6
with 1% or more terbium). The crust abundance of terbium is estimated as 1.2 mg/kg.[202] No terbium-dominant mineral has yet been found.[203]

Thalliums

[edit | edit source]
File:Thallium flame.png
The image shows the color of thallium in a natural gas burner. Credit: Ernest Z. {{fairuse}}
Pieces of very pure thallium are in a glass ampoule under argon. Credit: W. Oelen.{{free media}}

Thallium (I) ions are found geologically mostly in potassium-based ores. The radioisotope thallium-201 is the soluble chloride TlCl.

A mercury–thallium alloy, which forms a eutectic at 8.5% thallium, is reported to freeze at −60 °C, some 20 °C below the freezing point of mercury.

Thallium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

There is a green thallium line that shows up in arc spectra using "two to eight amperes at 120 volts, usually between ordinary arc carbons."[144]

Thoriums

[edit | edit source]
Thorium sample (99.9 % = 3N) is a thin sheet under argon in a glass ampoule, ca. 0.1 g. Credit: Alchemist-hp.{{free media}}

Thorium is a silvery, radioactive, metallic element. At room temperature and pressure, thorium crystallizes into a face-centered cubic lattice, where one thorium atom occupies each location of a black sphere in the diagram on the left.

Thorium can form alloys with many other metals. Addition of small proportions of thorium improves the mechanical strength of magnesium, and thorium-aluminium alloys have been considered as a way to store thorium in thorium nuclear reactors. Thorium forms eutectic mixtures with chromium and uranium, and it is completely miscible in both solid and liquid states with its lighter congener cerium.[204]

This diagram shows a face-centered cubic lattice. Credit: Daniel Mayer and DrBob.{{free media}}
Thorium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Def. a chemical element (symbol Th) with atomic number 90 is called thorium.

Tetravalent thorium compounds are usually colourless or yellow, like those of silver or lead, as the Th4+ ion has no 5f or 6d electrons.[205]

Monazites

[edit | edit source]
This monazite is a tabular crystal from Rostadheia, Iveland, Norway. Credit: Aangelo.{{free media}}
The primary source of the world's thorium is the rare-earth-and-thorium-phosphate mineral monazite. Credit: USGS.{{free media}}
Monazite gets its name from the Greek word "monazein", which means "to be alone", in allusion to its isolated crystals and their rarity when first found. Credit: Robert M. Lavinsky.{{free media}}

Monazite, a primarily reddish-brown phosphate mineral that contains rare-earth elements, with variability composition, is considered a group of minerals:[206]

  • monazite-(Ce), (Ce,La,Nd,Th)PO
    4
    (the most common member),
  • monazite-(La), (La,Ce,Nd)PO
    4
    ,
  • monazite-(Nd), (Nd,La,Ce)PO
    4
    ,
  • monazite-(Sm), (Sm,Gd,Ce,Th)PO
    4
    ,
  • monazite-(Pr), (Pr,Ce,Nd,Th)PO
    4
    .

(Ce,La,Nd,Th)PO
4
occurs usually in small isolated crystals has a hardness of 5.0 to 5.5 on the Mohs scale of mineral hardness and is relatively dense, about 4.6 to 5.7 g/cm3.

The primary source of the world's thorium is the rare-earth, and thorium, phosphate mineral monazite.

Silica (SiO
2
) is present in trace amounts, as is small amounts of uranium.

Due to the alpha decay of thorium and uranium, monazite contains a significant amount of helium, which can be extracted by heating.[207]

Umbozerites

[edit | edit source]
Umbozerite is from Karnasurt Mt., Lovozero Massif, Kola, Russia, size 4.2 cm. Credit: Weirdmeister.{{free media}}

The IMA-CNMNC approved mineral symbol is Ubz.[180]

Umbozerites have the chemical formula Na
3
Sr
4
Th[Si(O,OH
(3-4)
]
8
, IMA formula Na
3
Sr
4
ThSi
8
(O,OH)
24
, common impurities: Ti,Ce,Fe,U,Mn,Ca,Ba,K, and Crystal System: Amorphous.[208]

Environment: In ussingite veinlets cutting alkalic rocks, type locality: Umbozero (Lake Umba), Kola Peninsula, Russia, dark brown prismatic umbozerite masses in pegmatite rock, Metamict - Mineral originally crystalline, now amorphous due to radiation damage, Pseudo Tetragonal - Crystals show a tetragonal shape, Umbozerite is Radioactive as defined in 49 CFR 173.403, greater than 70 Bq / gram.[209]

Occurrence: In pneumatolytic-hydrothermal veins cutting alkalic rocks in the upper part of a differentiated alkalic massif, Crystal Data: Metamict; tetragonal after recrystallization[210]

Association: Ussingite, sphalerite, belovite, manganoan pectolite, lorenzenite, niobium-bearing minerals of the lomonosovite group.[210]

Distribution: Found on Mts. Karnasurt and Punkaruaiv, near Lake Umba, Lovozero massif, Kola Peninsula, Russia.[210]

Thuliums

[edit | edit source]
Thulium, sublimed-dendritic, high purity 99.99 % Tm/TREM. Credit: Alchemist-hp.{{free media}}
Thulium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Thulium dissolves readily in dilute sulfuric acid to form solutions containing the pale green Tm (III) ions, which exist as [Tm(OH
2
)
9
]3+
complexes:[211]

2Tm
(s)
+ 3H
2
SO
4(aq)
→ 2Tm3+
(aq)
+ 3SO2−
4 (aq)
+ 3H
2(aq)
This small piece of native tin is from the Badiko District, Bauchi State, Nigeria. Credit: Robert Lavinsky.{{free media}}
File:Native tin.jpg
This is native tin from a porphyry copper deposit in the Bingham Deposit, Utah, USA. Credit: Alison Roberts.{{fairuse}}

Native tin such as in the images on the right and left occurs in two crystal forms: α-Sn (cubic) and β-Sn (tetragonal).[19]

Purest tin 99,999 % = 5N, beta (left, white) and alpha (right, gray) allotropes. Credit: Alchemist-hp.{{free media}}
Tin spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

α-tin, the nonmetallic form or gray tin, is stable below 13.2 °C (55.8 °F) and is brittle. α-tin has a diamond cubic crystal structure, similar to diamond, silicon or germanium. α-tin has no metallic properties, because its atoms form a covalent structure in which electrons cannot move freely. α-tin is a dull-gray powdery material with no common uses other than specialized semiconductor applications.[212]

The α-β transformation temperature is 13.2 °C (55.8 °F), but impurities (e.g. Al, Zn, etc.) lower it well below 0 °C (32 °F). With the addition of antimony or bismuth the transformation might not occur at all, increasing durability.[213]

β–α transition of tin is at −40 °C.

β-tin, the metallic form or white tin, has Tetragonal crystal system, body-centered tetragonal (BCT structure) and is stable at and above room temperature and is malleable.

γ-tin and σ-tin exist at temperatures above 161 °C (322 °F)  and pressures above several GPa.[214]

Bronzes

[edit | edit source]

Bronze is an alloy consisting primarily of copper, commonly with about 12–12.5% tin and often with the addition of other metals (such as aluminum, manganese, nickel or zinc) and sometimes non-metals, such as phosphorus, or metalloids such as arsenic, or silicon.

Titaniums

[edit | edit source]
File:Iron-titanium phase diagram.png
This is an iron-titanium phase diagram. Credit: Hirokai.{{fairuse}}
This is an emission spectrum that covers the visible range: 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Titanium strips inside a glass jar are part of the Everest Element Set from Russia. Credit: RTC.{{free media}}
A titanium crystal bar, high purity 99.995 %, made by the iodide process at URALREDMET in the Soviet era. Credit: Alchemist-hp.{{free media}}

"Microbeam analysis of eclogites from the ultrahigh-pressure metamorphic belt in Dabieshan, China has revealed native titanium inclusions in garnets of coesite eclogite. The inclusions are about 10 μm in size, have a submetallic luster from the thin oxidation film on the surface, and are brown under reflected light."[215]

Titanium is a dimorphic allotrope that "undergoes a phase transformation (hcp to bcc) at 882 °C [5]."[216]

As the phase diagram on the left indicates, there is a miscibility gap between bcc iron (α-Fe) and hcp (α-Ti) up to about 800°C.

Titanium (Ti) has green emission lines at 521.97, 522.268, 522.413, 524.729, and 526.596 nm as observed in solar limb faculae.[217]

Titanium (Ti II) has an absorption band, 391.346-441.108 nm, with an excitation potential range of 0.60-3.08 eV.[84]

Titanium has two emission lines at 456.3757 and 457.1971 nm from Ti II.[218]

Titanium can be alloyed with iron, aluminium, zirconium, nickel, vanadium, copper, and molybdenum.

Osbornites

[edit | edit source]

Osbornite is a very rare natural form of titanium nitride (TiN), found almost exclusively in meteorites.[219][220]

Tungstens

[edit | edit source]
File:Native tungsten Luna 24 landing site.jpg
The small, bright crystalline mass on the right of this electron micrograph is native tungsten. Credit: Andrei V. Mokhov.{{fairuse}}
File:Fe-W-phase-diagram.jpg
This is an iron-tungsten phase diagram. Credit: Satyendra.{{fairuse}}
Tungsten rods are shown with evaporated crystals, partially oxidized with colorful tarnish. Credit: Alchemist-hp.{{free media}}
Tungsten spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

In the scanning electron micrograph on the right is a bright grain, or crystalline mass, of native tungsten. The sample is a fragment of lunar silicate glass from the Luna 24 landing site, Mare Crisium, The Moon. The fragment is bright in backscattered electrons.

The iron-tungsten phase diagram on the left shows that the bcc phase of tungsten (α-W) occurs from lower temperatures on up to the melting temperature.

Tungsten is usually alloyed with nickel, iron, or cobalt to form heavy alloys,

Tungsten carbide (chemical formula: WC) is a chemical compound (specifically, a carbide) containing equal parts of tungsten and carbon atoms.

Uraniums

[edit | edit source]
Uranium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
A billet of highly enriched uranium was recovered from scrap processed at the Y-12 National Security Complex Plant. Credit: Choihei.{{free media}}

Uranium "not only exists in the forms of tetravalent and hexavalent uranium oxides, but also occurs in the form of native uranium [from the hydrothermal Guidong and Zhuguang uranium deposits of the middle Nanling metallogenic belt, Southern China]."[221]

Depleted uranium (DU) is alloyed with 1–2% other elements, such as titanium or molybdenum.[222] UCo is a superconductor at 1.70°K.[223]

UMnGe (Pnma, a = 686.12(9), b = 425.49(6) and c = 736.5(1) pm) adopts the orthorhombic structure of TiNiSi and U
2
Mn
3
Ge
(P63/mmc, a = 524.3(2) and c = 799.2(3) pm) possesses the hexagonal Mg
2
Cu
3
Si
-type structure (ordered variant of the hexagonal Laves phase MgZn
2
).[224]

Vanadiums

[edit | edit source]
File:Native vanadium crystals.png
In this backscattered electron micrograph on the left, the native vanadium crystals have been colorized in red. Credit: MikhailI Ostrooumov and Yuri Taran.{{fairuse}}
File:Fe-V Phase Diagram.gif
This Fe-V phase diagram shows which phases are to be expected at equilibrium for different combinations of vanadium content and temperature. Credit: Computational Thermodynamics Inc.{{fairuse}}
Vanadium pieces are shown. Credit: W. Oelen.{{free media}}
Oxidation states of vanadium are shown from left +2 (lilac), +3 (green), +4 (blue) and +5 (yellow). Credit: Steffen Kristensen. {{free media}}

"[N]ative vanadium [occurs] in natural fumarolic incrustations and in the mineral assemblage precipitated in silica tubes inserted into high-temperature (750-830°C) fumaroles of Colima volcano – the most active volcano of Mexico, and one of the most active in the Americas. [...] The new mineral and its name (“vanadium”) have been approved by the Commission on New Minerals, Nomenclature and Classification of the International Mineralogical Association (Williams et al., 2013; IMA # 2012- 021a). The holotype material has been deposited in the Geological Museum of National Mexican University (New mineral collection of Mexican Mineralogical Society with cataloged under FIM 12/01)."[225]

In the image on the right, the backscattered electron micrograph on the left side, has the native vanadium crystals colorized in red. The energy dispersive X-ray spectroscopy (EDS) spectrum on the right shows the vanadium peaks plus small amounts of Fe and S.[225]

As the phase diagram on the left indicates vanadium is bcc down to lower temperatures from its melting point.

This is an emission spectrum that covers the visible range: 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The chemistry of vanadium is noteworthy for the accessibility of the four adjacent oxidation states 2-5. In aqueous solution the colours are lilac V2+(aq), green V3+(aq), blue VO2+(aq) and, at high pH, yellow VO42-.

Vanadium (V II) has an absorption band, 392.973-403.678 nm, with an excitation potential range of 1.07-1.81 eV.[84]

Xenons

[edit | edit source]

Ytterbiums

[edit | edit source]
Ytterbium metal solid is 0.5 x 1 cm. Credit: Jurii.{{free media}}
Euxenite crystal aggregation (11 cm) is from Vegusdal, Norway. Credit: Aangelo.{{free media}}

Natural ytterbium is a mixture of seven stable isotopes, which altogether are present at concentrations of 0.3 parts per million.

Ytterbium has three allotropes: alpha, beta and gamma; their transformation temperatures are −13 °C and 795 °C,[50] although the exact transformation temperature depends on the pressure and stress.[226] The beta allotrope (6.966 g/cm3) exists at room temperature, and it has a face-centered cubic crystal structure. The high-temperature gamma allotrope (6.57 g/cm3) has a body-centered cubic crystalline structure.[50] The alpha allotrope (6.903 g/cm3) has a hexagonal crystalline structure and is stable at low temperatures.[131] The beta allotrope has a metallic electrical conductivity at normal atmospheric pressure, but it becomes a semiconductor when exposed to a pressure of about 16,000 atmospheres (1.6 GPa). Its electrical resistivity increases ten times upon compression to 39,000 atmospheres (3.9 GPa), but then drops to about 10% of its room-temperature resistivity at about 40,000 atm (4.0 GPa).[50][68]

Ytterbium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}

The alpha allotrope is diamagnetic.[226]

Ytterbium is paramagnetic at temperatures above 1.0 K.[227]

Natural ytterbium is composed of seven stable isotopes: 168Yb, 170Yb, 171Yb, 172Yb, 173Yb, 174Yb, and 176Yb, with 174Yb being the most common, at 31.8% of the natural abundance). 27 radioisotopes have been observed, with the most stable ones being 169Yb with a half-life of 32.0 days, 175Yb with a half-life of 4.18 days, and 166Yb with a half-life of 56.7 hours. All of the remaining radioactive isotopes have half-lives that are less than two hours, and most of these have half-lives under 20 minutes. Ytterbium also has 12 meta states, with the most stable being 169mYb (t1/2 46 seconds).[228][89]

The isotopes of ytterbium range in atomic weight from 147.9674 atomic mass unit (u) for 148Yb to 180.9562 u for 181Yb. The primary decay mode of ytterbium isotopes lighter than the most abundant stable isotope, 174Yb, is electron capture, and the primary decay mode for those heavier than 174Yb is beta decay. The primary decay products of ytterbium isotopes lighter than 174Yb are thulium isotopes, and the primary decay products of ytterbium isotopes with heavier than 174Yb are lutetium isotopes.[228][89]

It occurs in the minerals monazite, euxenite, and xenotime.

Ytterbium Oxides

[edit | edit source]

The +2 oxidation state occurs only in solid compounds and reacts in some ways similarly to the alkaline earth metal compounds; for example, ytterbium(II) oxide (YbO) shows the same structure as calcium oxide (CaO).[131]

Yttriums

[edit | edit source]
Yttrium, sublimed-dendritic, high purity 99.99 % Y/TREM, as well as an argon arc remelted 1 cm3 yttrium cube for comparison, purity 99.9 %. Credit: Alchemist-hp.{{free media}}
Yttrium spectrum is 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Xenotime crystals contain yttrium. Credit: Tom Epaminondas (mineral collector) / Eurico Zimbres (FGEL-UERJ).{{free media}}

Yttrium (Y II) has an absorption band from 395.035 to 439.802 nm, with an excitation potential range of 0.10-0.13 eV.[84]

The orange system, in orange astronomy is a number of emission lines very close together forming a band in the orange portion of the visible spectrum. These lines are usually associated with particular molecular species, including ScO, YO, and TiO.[229]

Small amounts of yttrium (0.1 to 0.2%) have been used to reduce the grain sizes of chromium, molybdenum, titanium, and zirconium.[230] Yttrium is used to increase the strength of aluminium and magnesium alloys.[231] The addition of yttrium to alloys generally improves workability, adds resistance to high-temperature recrystallization, and significantly enhances resistance to high-temperature oxidation.[232]

Yttrium nitrides

[edit | edit source]

Yttrium nitride (YN) is formed when the metal is heated to 1000 °C in nitrogen.[232]

Zincs

[edit | edit source]
Zinc emission spectrum is for 400 nm - 700 nm. Credit: McZusatz.{{free media}}
Pure zinc metal wedge is shown. Credit: United States Geological Survey.{{free media}}
The image shows the color of zinc in a natural gas burner. Credit: Chemicalinterest.{{free media}}
Zinc, purity 99.995 %, left: a crystaline fragment of an ingot, right: sublimed-dendritic, and a 1 cm3 zinc cube for comparison. Credit: Alchemist-hp.{{free media}}

"Satellite images taken over the past several decades show the dramatic disappearance of ice, including on the island’s inland areas, where the ice fields can in places be up to three and a half kilometers deep."[51]

"Along with uranium, zinc, iron ore, copper and gold, Greenland’s ancient rocks also harbor large quantities of those minerals known as “rare earth,” among them lanthanum, cerium, neodymium, praesodymium, terbium and yttrium."[51]

Metals long known to form binary alloys with zinc are aluminium, antimony, bismuth, gold, iron, lead, mercury, silver, tin, magnesium, cobalt, nickel, tellurium, and sodium.[233]

Although neither zinc nor zirconium is ferromagnetic, their alloy ZrZn
2
exhibits ferromagnetism below 35 K.[234]

Brasses

[edit | edit source]
Copper-zinc (Cu-Zn) binary phase diagram is for brass. Credit: Cdang.{{free media}}

The earliest brasses may have been natural alloys made by smelting zinc-rich copper ores.[235]

The compositions of these early "brass" objects are highly variable and most have zinc contents of between 5% and 15% wt which is lower than in brass produced by cementation.[235]

Alpha-brass is Cu
3
Zn
.[236]

Zincites

[edit | edit source]
This giant chunk of a crystal cluster is enormous. Credit: cobalt123.{{free media}}

Zincite has the formula ZnO.[237]

  1. Colour: Red, orange, yellow, white; rarely green.[237]
  2. Lustre: Sub-Vitreous, Resinous, Waxy, Greasy, Silky, Dull, Earthy.[237]
  3. Crystal System: Hexagonal.[237]

Zinc selenides

[edit | edit source]

"ZnSe appears as an attractive material to blue and near UV optoelectronics."[238]

Zirconiums

[edit | edit source]
Purest zirconium 99.97%, two samples of crystal bar showing different surface textures, made by crystal bar process, as well as a highly pure (99,95 % = 3N5) 1 cm3 zirconium cube for comparison. Credit: Alchemist-hp.{{free media}}
File:Fe-Zr phase diagram.gif
This is a binary phase diagram of the iron-zirconum system. Credit: D. Arias and J.P. Abriata.{{fairuse}}
Zirconium spectrum is for 400 nm - 700 nm. Credit: McZusatz.{{free media}}

Zirconium is a lustrous, greyish-white, soft, ductile, malleable metal that is solid at room temperature, though it is hard and brittle at lesser purities.[53][239]

Zirconium is highly resistant to corrosion by alkalis, acids, salt water and other agents.[231] However, it will dissolve in hydrochloric and sulfuric acid, especially when fluorine is present.[240]

Alloys with zinc are magnetic at less than 35 K.[231]

The melting point of zirconium is 1855 °C (3371 °F), and the boiling point is 4371 °C (7900 °F).[241] Zirconium has an electronegativity of 1.33 on the Pauling scale for the elements within the d-block with known electronegativities, zirconium has the fifth lowest electronegativity after hafnium, yttrium, lanthanum, and actinium.[242]

At room temperature zirconium exhibits a hexagonally close-packed crystal structure, α-Zr, which changes to β-Zr, a body-centered cubic crystal structure, at 863 °C, β-phase until the melting point.[243]

As the Fe-Zr phase diagram on the left demonstrates, zirconium has a hcp structure (α-Zr) at lower temperatures, including room temperature, and a bcc structure (β-Zr) at higher temperatures up to melting.

"Zirconium isotopic abundances [may be] determined from ZrO bandheads near 6925 Å via synthetic spectra for a sample of S stars."[244]

Zirconium (Zr II) has an absorption band, 395.824-415.624 nm, with an excitation potential of 0.52-0.75 eV.[84]

The mineral zircon is the most important source of zirconium.

Naturally occurring zirconium is composed of five isotopes:

  1. 90Zr is the most common, making up 51.45% of all zirconium,
  2. 91Zr,
  3. 92Zr and
  4. 94Zr are stable, although 94Zr is predicted to undergo double beta decay (not observed experimentally) with a half-life of more than 1.10×1017 years,
  5. 96Zr has a half-life of 2.4×1019 years, is the longest-lived radioisotope of zirconium and is the least common, comprising only 2.80% of zirconium.[245]

Twenty-eight artificial isotopes of zirconium have been synthesized, ranging in atomic mass from 78 to 110.

  1. 88Zr, decays by electron capture,
  2. 93Zr is the longest-lived artificial isotope, with a half-life of 1.53×106 years,
  3. 110Zr, the heaviest isotope of zirconium, is the most radioactive, with an estimated half-life of 30 milliseconds.[245]

Radioactive isotopes at or above mass number 93 decay by electron emission, whereas those at or below 89 usually decay by positron emission.

Five isotopes of zirconium also exist as metastable isomers:

  1. 83mZr,
  2. 85mZr,
  3. 89mZr, is the longest lived with a half-life of 4.161 minutes,
  4. 90m1Zr,
  5. 90m2Zr has the shortest half-life at 131 nanoseconds, and
  6. 91mZr.[245]

88Zr: "When irradiated with low-energy neutrons from a nuclear reactor, each atom of zirconium-88 had a high probability of absorbing a neutron into its nucleus, causing the element to transform into another isotope, zirconium-89. The reaction was about 85,000 times as likely to occur as predicted."[246]

"88Zr has a thermal neutron capture cross-section of 861,000 ± 69,000 barns (1σ uncertainty), which is five orders of magnitude larger than the theoretically predicted value of 10 barns2."[247]

"Only one other isotope, xenon-135, is known to be better at capturing neutrons. Previously studied versions of zirconium are much more reluctant to take on another neutron, with absorption probabilities about a millionth that of zirconium-88, or less."[246]

"Isotopes with a high neutron capture probability can be used to control nuclear reactors by sopping up loose neutrons, slowing the rate of reactions."[246]

Hypotheses

[edit | edit source]
  1. The use of satellites should provide ten times the information as sounding rockets or balloons.

A control group for a radiation satellite would contain

  1. a radiation astronomy telescope,
  2. a two-way communication system,
  3. a positional locator,
  4. an orientation propulsion system, and
  5. power supplies and energy sources for all components.

A control group for radiation astronomy satellites may include an ideal or rigorously stable orbit so that the satellite observes the radiation at or to a much higher resolution than an Earth-based ground-level observatory is capable of.

See also

[edit | edit source]

References

[edit | edit source]
  1. Christoph Seidler; translated by Anne-Marie de Grazia (19 June 2014). "Earth's weakening magnetic field". Q-Mag.org. Retrieved 2014-10-21.
  2. Muhammad moolla (15 May 2009). Topic:Mining. San Francisco, California: Wikimedia Foundation, Inc. http://en.wikiversity.org/wiki/Wikiversity:RFD#Dominant_group. Retrieved 2016-05-05. 
  3. Callister, W.D. "Materials Science and Engineering: An Introduction" 2007, 7th edition, John Wiley and Sons, Inc. New York, Section 4.3 and Chapter 9.
  4. Bluelion~enwiktionary (7 May 2003). alloy. San Francisco, California: Wikimedia Foundation, Inc. https://en.wiktionary.org/wiki/aloy. Retrieved 30 June 2022. 
  5. Widsith (1 May 2012). alloy. San Francisco, California: Wikimedia Foundation, Inc. https://en.wiktionary.org/wiki/alloy. Retrieved 30 June 2022. 
  6. Jtle515 (3 March 2012). admixture. San Francisco, California: Wikimedia Foundation, Inc. https://en.wiktionary.org/wiki/admixture. Retrieved 30 June 2022. 
  7. Paul G (8 November 2005). compound. San Francisco, California: Wikimedia Foundation, Inc. https://en.wiktionary.org/wiki/compound. Retrieved 30 June 2022. 
  8. 8.0 8.1 8.2 Quercus solaris (12 February 2022). compound. San Francisco, California: Wikimedia Foundation, Inc. https://en.wiktionary.org/wiki/compound. Retrieved 30 June 2022. 
  9. DCDuring (27 November 2010). compound. San Francisco, California: Wikimedia Foundation, Inc. https://en.wiktionary.org/wiki/compound. Retrieved 30 June 2022. 
  10. DavidL2 (12 October 2004). compound. San Francisco, California: Wikimedia Foundation, Inc. https://en.wiktionary.org/wiki/compound. Retrieved 30 June 2022. 
  11. Verhoeven, John D. (2007). Steel Metallurgy for the Non-metallurgist. ASM International. p. 56. https://web.archive.org/web/20160505065853/https://books.google.com/books?id=brpx-LtdCLYC&pg=PA56. 
  12. Mills, Adelbert Phillo (1922) Materials of Construction: Their Manufacture and Properties, John Wiley & sons, inc, originally published by the University of Wisconsin, Madison
  13. 13.0 13.1 Vesselin M. Dekov, Vasil Arnaudov, Frans Munnik, Tanya B. Boycheva, and Saverio Fiore (August 2009). "Native aluminum: Does it exist?". American Mineralogist 94 (8-9): 1283-6. doi:10.2138/am.2009.3236. http://rruff.info/uploads/AM94_1283.pdf. Retrieved 2015-08-28. 
  14. Barthelmy, D. "Aluminum Mineral Data". Mineralogy Database. Retrieved 9 July 2008. {{cite web}}: |archive-date= requires |archive-url= (help)
  15. Chen, Z.; Huang, Chi-Yue; Zhao, Meixun; Yan, Wen; Chien, Chih-Wei; Chen, Muhong; Yang, Huaping; Machiyama, Hideaki et al. (2011). "Characteristics and possible origin of native aluminum in cold seep sediments from the northeastern South China Sea". Journal of Asian Earth Sciences 40 (1): 363–370. doi:10.1016/j.jseaes.2010.06.006. 
  16. aluminide. San Francisco, California: Wikimedia Foundation, Inc. 31 March 2014. https://en.wiktionary.org/wiki/aluminide. Retrieved 2015-02-22. 
  17. D. Stammose and J.-M. Dolo (1990). "Sorption of americium at trace levels on a clay mineral". Radiochimica Acta 51: 189-93. https://books.google.com/books?id=AP3wCAAAQBAJ&pg=PA202&lpg=PA202&source=bl&ots=HSbrCjCBzW&sig=TCbSmutAUaFoz9bY4OEShSLPcSw&hl=en&sa=X&ved=0CDsQ6AEwBWoVChMIz8GJ9P_6yAIVRuYmCh24kA0R#v=onepage&f=false. Retrieved 2015-11-05. 
  18. M. R. Melloch; J. M. Woodall; E. S. Harmon; N. Otsuka; Fred H. Pollak; D. D. Nolte; R. M. Feenstra; M. A. Lutz (1995). "Low-temperature grown III-V materials". Annual Review of Materials Science 25 (1): 547-600. doi:10.1146/annurev.ms.25.080195.002555. http://www.annualreviews.org/doi/pdf/10.1146/annurev.ms.25.080195.002555. Retrieved 2013-08-29. 
  19. 19.0 19.1 19.2 19.3 19.4 19.5 19.6 Willard Lincoln Roberts; George Robert Rapp Jr.; Julius Weber (1974). Encyclopedia of Minerals. New York, New York, USA: Van Nostrand Reinhold Company. pp. 121-2. 
  20. jolyon (6 November 2015). "The Mineralogy of Curium". Hudson Institute of Mineralogy. Retrieved 2015-11-05.
  21. A. P. Meshik (November 2005). "The Workings of an Ancient Nuclear Reactor". Scientific American. http://www.sciam.com/article.cfm?id=ancient-nuclear-reactor. 
  22. F. Gauthier-Lafaye; P. Holliger; P.-L. Blanc (1996). "Natural fission reactors in the Franceville Basin, Gabon: a review of the conditions and results of a "critical event" in a geologic system". Geochimica et Cosmochimica Acta 60 (25): 4831–52. doi:10.1016/S0016-7037(96)00245-1. 
  23. http://www.matthey.ch/index.php?id=45&L=1 Copper beryllium and nickel beryllium datasheets.
  24. "Beralcast – Beryllium Aluminum Alloys". IBC Advanced Alloys. Retrieved 2015-07-22. {{cite web}}: |archive-date= requires |archive-url= (help)
  25. 25.0 25.1 J. M. Jacobs & M. M. Dworetsky (7 October 1982). "Bismuth abundance anomaly in a Hg—Mn star". Nature 299: 535–536. https://www.nature.com/articles/299535a0. Retrieved 20 June 2022. 
  26. "Q & A: Where does the element Boron come from?". Retrieved 2011-12-04. {{cite web}}: |archive-date= requires |archive-url= (help)
  27. "Boron". Britannica encyclopedia. Retrieved 4 August 2020. {{cite web}}: |archive-date= requires |archive-url= (help)
  28. Thomas K. Gaisser (1990). Cosmic Rays and Particle Physics. Cambridge University Press. pp. 279. http://books.google.com/books?hl=en&lr=&id=qJ7Z6oIMqeUC&oi=fnd&pg=PR15&ots=IxjwLxBwXu&sig=voHKIYstBlBYla4jcbur_b-Zwxs. Retrieved 2014-01-11. 
  29. 29.0 29.1 Qingsongite on Mindat.org
  30. Qingsongite: New Mineral from Tibet Hard as Diamond. sciencenews.org. August 5, 2013
  31. Pittalwala, Iqbal, International Research Team Discovers New Mineral, UCR Today, Aug. 2, 2013
  32. Griggs, Jessica (2014-05-13). "Diamond no longer nature's hardest material". New Scientist. Retrieved 2018-01-12.
  33. 33.0 33.1 33.2 33.3 Scoullos, Michael J.; Vonkeman, Gerrit H.; Thornton, Iain; Makuch, Zen (2001). Mercury, Cadmium, Lead: Handbook for Sustainable Heavy Metals Policy and Regulation. Springer. https://books.google.com/books?id=9yzN-QGag_8C. 
  34. Jennings, Thomas C. (2005). "Cadmium Environmental Concerns". PVC handbook. Hanser Verlag. p. 149. https://books.google.com/books?id=YUkJNI9QYsUC&pg=PA149. 
  35. Brady, George Stuart; Brady, George S.; Clauser, Henry R.; Vaccari, John A. (2002). Materials handbook: an encyclopedia for managers, technical professionals, purchasing and production managers, technicians, and supervisors. McGraw-Hill Professional. p. 425. ISBN 978-0-07-136076-0. https://books.google.com/books?id=vIhvSQLhhMEC&pg=PA425. 
  36. 36.0 36.1 36.2 36.3 Haire, Richard G. (2006). "Californium". In Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Springer Science+Business Media. ISBN 978-1-4020-3555-5.
  37. Szwacki, Nevill Gonzalez; Szwacka, Teresa (2010). Basic Elements of Crystallography. Pan Stanford. ISBN 978-981-4241-59-5.
  38. K. J. McCarthy; A. Baciero; B. Zurro; TJ-II Team (12 June 2000). Impurity Behaviour Studies in the TJ-II Stellarator, In: 27th EPS Conference on Contr. Fusion and Plasma Phys.. 24B. Budapest: ECA. pp. 1244-7. http://crpppc42.epfl.ch/Buda/pdf/p3_116.pdf. Retrieved 20 January 2013. 
  39. 39.0 39.1 39.2 39.3 Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. ISBN 978-0-08-037941-8.
  40. Klaus Reinhardt and Herwig Winkler in "Cerium Mischmetal, Cerium Alloys, and Cerium Compounds" in Ullmann's Encyclopedia of Industrial Chemistry 2000, Wiley-VCH, Weinheim. doi:10.1002/14356007.a06_139
  41. Sims, Zachary (2016). "Cerium-Based, Intermetallic-Strengthened Aluminum Casting Alloy: High-Volume Co-product Development". JOM 68 (7): 1940–1947. doi:10.1007/s11837-016-1943-9. https://www.osti.gov/biblio/1257369. 
  42. Wen-Ji Bai, Mei-Fu Zhou, and Paul T. Robinson (August 1993). "Possibly diamond-bearing mantle peridotites and podiform chromitites in the Luobusa and Donqiao ophiolites, Tibet". Canadian Journal of Earth Sciences 30 (8): 1650-9. doi:10.1139/e93-143. http://www.nrcresearchpress.com/doi/abs/10.1139/e93-143. Retrieved 2015-08-19. 
  43. Subramanian, S. "Making New Elements Doesn't Pay. Just Ask This Berkeley Scientist". Bloomberg Businessweek. Retrieved 2020-01-18.
  44. A.B. Johnson Jr.; B. Francis (1 January 1980). Durability of metals from archaeological objects, metal meteorites, and native metals. PNL-3198. Richland, Washington USA: Battelle Pacific Northwest Laboratories, Department of Energy. doi:10.2172/5406419. http://www.osti.gov/scitech/biblio/5406419. Retrieved 2014-10-28. 
  45. I.M. Kettles; A.N. Rencz; S.D. Bauke (April 2000). "Integrating Landsat, Geologic, and Airborne Gamma Ray Data as an Aid to Surficial Geology Mapping and Mineral Exploration in the Manitouwadge Area, Ontario". Photogrammetric Engineering & Remote Sensing 66 (4): 437-45. http://asprs.org/a/publications/pers/2000journal/april/2000_apr_437-445.pdf. Retrieved 2014-10-28. 
  46. jolyon (6 November 2015). "The Mineralogy of Curium". Hudson Institute of Mineralogy. Retrieved 2015-11-05.
  47. 47.0 47.1 47.2 Moody, Ken (2013-11-30). "Synthesis of Superheavy Elements". In Schädel, Matthias; Shaughnessy, Dawn. The Chemistry of Superheavy Elements (2nd ed.). Springer Science & Business Media. pp. 24–8. ISBN 9783642374661. 
  48. Nagame, Y.; Kratz, J. V.; Schädel, M. (2016). "Chemical properties of rutherfordium (Rf) and dubnium (Db) in the aqueous phase". EPJ Web of Conferences 131: 07007. doi:10.1051/epjconf/201613107007. https://jopss.jaea.go.jp/pdfdata/BB2016-0022.pdf. 
  49. D Chatterjee and K N R Taylor (1972). "Magnetic and structural properties of the neodymium-dysprosium alloy system". Journal of Physics F: Metal Physics 2 (1): 151. https://iopscience.iop.org/article/10.1088/0305-4608/2/1/020/pdf. Retrieved 18 June 2022. 
  50. 50.0 50.1 50.2 50.3 50.4 Hammond, C. R. (2000). The Elements, in Handbook of Chemistry and Physics (81st ed.). CRC press. 
  51. 51.0 51.1 51.2 51.3 Silvia von der Weiden (21 March 2012). As Greenland's Glaciers Recede, A Rush On The Riches Buried Below. WorldCrunch. http://www.worldcrunch.com/business-finance/as-greenland-s-glaciers-recede-a-rush-on-the-riches-buried-below/c2s4915/. Retrieved 20 September 2014. 
  52. Wei Wang, Hui Zou, Shuizhou Cai (18 February 2019). "The Oxidation and Combustion Properties of Gas Atomized Aluminum− Boron− Europium Alloy Powders". Propellants, Explosives, Pyrotechnics 44 (6): 725-732. doi:10.1002/prep.201800223. https://onlinelibrary.wiley.com/doi/abs/10.1002/prep.201800223. Retrieved 18 June 2022. 
  53. 53.0 53.1 53.2 53.3 Emsley, John (2001). Nature's Building Blocks. Oxford: Oxford University Press. pp. 262–266. https://books.google.com/books?id=j-Xu07p3cKwC&pg=PA265. 
  54. Lavrukhina, Avgusta Konstantinovna; Pozdnyakov, Aleksandr Aleksandrovich (1970). Analytical Chemistry of Technetium, Promethium, Astatine, and Francium. Translated by R. Kondor. Ann Arbor–Humphrey Science Publishers. p. 269. ISBN 978-0-250-39923-9.
  55. Gschneidner, Karl Jr; Gibson, Kerry (7 December 2001). "Magnetic refrigerator successfully tested". Ames Laboratory. Retrieved 17 December 2006. {{cite web}}: |archive-date= requires |archive-url= (help)
  56. Gschneidner, K.; Pecharsky, V.; Tsokol, A. (2005). "Recent Developments in Magnetocaloric Materials". Reports on Progress in Physics 68 (6): 1479. doi:10.1088/0034-4885/68/6/R04. https://web.archive.org/web/20141109081936/http://www.teknik.uu.se/ftf/education/magnetmatr/Projektreferenser/MCE_Reports_Progress05.pdf. 
  57. 57.0 57.1 Alan M. Shiller (June 1998). "Dissolved gallium in the Atlantic Ocean". Marine Chemistry 61 (1-2): 87-99. doi:10.1016/S0304-4203(98)00009-7. http://www.sciencedirect.com/science/article/pii/S0304420398000097. Retrieved 2014-10-29. 
  58. A. Robertson Jr; T.H. Chiu; W.T. Tsang; J.E. Cunningham (1987). "RHEED Intensity Oscillation Studies of the Kinetics of GaAs Deposition During Chemical Beam Epitaxy (CBE)". MRS Proceedings 102: 17-23. doi:http://dx.doi.org/10.1557/PROC-102-17. http://journals.cambridge.org/abstract_S1946427400538469. Retrieved 2012-07-17. 
  59. R Nave (17 August 2015). "Germanite". Washington, DS USA: Smithsonian Museum of Natural History. Retrieved 2015-08-17.
  60. Hebda, John (2001). "Niobium alloys and high Temperature Applications" (PDF). CBMM. Retrieved 2008-09-04. {{cite web}}: |archive-date= requires |archive-url= (help)
  61. Maslenkov, S. B.; Burova, N. N.; Khangulov, V. V. (1980). "Effect of hafnium on the structure and properties of nickel alloys". Metal Science and Heat Treatment 22 (4): 283–285. doi:10.1007/BF00779883. 
  62. Beglov, V. M.; Pisarev, B. K.; Reznikova, G. G. (1992). "Effect of boron and hafnium on the corrosion resistance of high-temperature nickel alloys". Metal Science and Heat Treatment 34 (4): 251–254. doi:10.1007/BF00702544. 
  63. Voitovich, R. F.; Golovko, É. I. (1975). "Oxidation of hafnium alloys with nickel". Metal Science and Heat Treatment 17 (3): 207–209. doi:10.1007/BF00663680. 
  64. A. N. Bushuev, O. V. El’kin, I. V. Tolstobrova, A. V. Sazanov, and D. A. Kondrat’ev (2018). "Preparation of a Nickel–Holmium Alloy Coating in an Equimolar HoCl
    3
    -Containing NaCl–KCl Melt"
    . Russian Metallurgy (Metally) 2018 (8): 771-778. https://www.researchgate.net/profile/Andrey-Bushuev-2/publication/329899928_Preparation_of_a_Nickel-Holmium_Alloy_Coating_in_an_Equimolar_HoCl3-Containing_NaCl-KCl_Melt/links/5dd4e7bf299bf11ec8629a72/Preparation-of-a-Nickel-Holmium-Alloy-Coating-in-an-Equimolar-HoCl3-Containing-NaCl-KCl-Melt.pdf. Retrieved 18 June 2022.
     
  65. Surmann, P; Zeyat, H (Nov 2005). "Voltammetric analysis using a self-renewable non-mercury electrode". Analytical and Bioanalytical Chemistry 383 (6): 1009–13. doi:10.1007/s00216-005-0069-7. PMID 16228199. 
  66. Nilson F. Botelho; Guy Roger; Ferdinand d'Yvoire; Yves Moẽlo; Marcel Volfinger (1994). "Yanomamite, InAsO4.2H2O, a new indium mineral from topaz-bearing greisen in the Gohiás Tin Province, Brazil". European Journal of Mineralogy 6: 245-54. https://rruff-2.geo.arizona.edu/uploads/EJM6_245.pdf. Retrieved 2015-02-22. 
  67. Hunt, L. B. (1987). "A History of Iridium". Platinum Metals Review 31 (1): 32–41. http://www.platinummetalsreview.com/pdf/pmr-v31-i1-032-041.pdf. 
  68. 68.0 68.1 Emsley, J. (2003). "Iridium". Nature's Building Blocks: An A–Z Guide to the Elements. Oxford, England, UK: Oxford University Press. pp. 201–204]. https://archive.org/details/naturesbuildingb0000emsl/page/201. 
  69. Reed-Hill R, Abbaschian R (1991). Physical Metallurgy Principles (3rd ed.). Boston: PWS-Kent Publishing. 
  70. Gove PB, ed (1963). Webster's Seventh New Collegiate Dictionary. Springfield, Massachusetts, USA: G & C Merriam Company. p. 58. 
  71. John G. Phillips; Sumner P. Davis; Bo Lindgren; Walter J. Balfour (December 1987). "The near-infrared spectrum of the FeH molecule". The Astrophysical Journal Supplement Series 65 (12): 721-78. doi:10.1086/191241. 
  72. DE Fawzy; NH Youssef; O. Engvold (May 1998). "Identification of FeH molecular lines in the spectrum of a sunspot umbra". Astronomy and Astrophysics Supplement 129 (5): 435-43. doi:10.1051/aas:1998196. http://aas.aanda.org/articles/aas/abs/1998/09/h0667/h0667.html. Retrieved 2012-02-18. 
  73. Lide, D. R., ed. (2005). CRC Handbook of Chemistry and Physics (86th ed.). Boca Raton (FL): CRC Press. ISBN 0-8493-0486-5.
  74. Uchida, H. (1999). "Hydrogen solubility in rare earth based hydrogen storage alloys". International Journal of Hydrogen Energy 24 (9): 871–877. doi:10.1016/S0360-3199(98)00161-X. 
  75. 75.0 75.1 Sato, T. K.; Asai, M.; Borschevsky, A.; Stora, T.; Sato, N.; Kaneya, Y.; Tsukada, K.; Düllman, Ch. E. et al. (9 April 2015). "Measurement of the first ionization potential of lawrencium, element 103". Nature 520 (7546): 209–11. doi:10.1038/nature14342. PMID 25855457. http://cds.cern.ch/record/2008656/files/TKSato-Lr-IP_prep_nature.pdf. 
  76. Gunther, Matthew (9 April 2015). "Lawrencium experiment could shake up periodic table". RSC Chemistry World. Retrieved 21 September 2015.
  77. 77.0 77.1 Haire, R. G. (11 October 2007). "Insights into the bonding and electronic nature of heavy element materials". Journal of Alloys and Compounds 444–5: 63–71. doi:10.1016/j.jallcom.2007.01.103. https://zenodo.org/record/1259091. 
  78. S.K. Peneva, K.D. Djuneva and E.A. Tsukeva (2 May 1981). "RHEED study of the initial stages of crystallization and oxidation of lead and tin". Journal of Crystal Growth 53 (2): 382-396. doi:10.1016/0022-0248(81)90088-9. http://www.sciencedirect.com/science/article/pii/0022024881900889. Retrieved 2017-12-13. 
  79. "The Eight Metals"
  80. Social, Cultural, and Economic History of Himachal Pradesh. Manjit Singh Ahluwalia. Indus Publishing. 1998 p. 163.
  81. स्वर्ण रूप्यं ताम्रं च रंग यशदमेव च। शीसं लौहं रसश्चेति धातवोऽष्टौ प्रकीर्तिता:। Here rasa can be taken as either mercury or brass.
  82. 82.0 82.1 Isaac Babbitt, http://pdfpiw.uspto.gov/.piw?Docid=00001252&idkey=NONE&homeurl=http%3A%252F%252Fpatft.uspto.gov%252Fnetahtml%252FPTO%252Fpatimg.htm "Mode of making boxes for axles and gudgeons," U.S. patent no. 1,252 (issued: July 17, 1839).
  83. Jeremy R. King; Constantine P. Deliyannis; Merchant Boesgaard (April 1, 1997). "The 9Be Abundances of α Centauri A and B and the Sun: Implications for Stellar Evolution and Mixing". The Astrophysical Journal 478 (2): 778. http://iopscience.iop.org/0004-637X/478/2/778/pdf/0004-637X_478_2_778.pdf. Retrieved 2012-07-11. 
  84. 84.00 84.01 84.02 84.03 84.04 84.05 84.06 84.07 84.08 84.09 84.10 84.11 Kozo Sadakane; Minoru Ueta (August 1989). "Abundance Analysis of Sirius in the Blue-Violet Region". Publications of the Astronomical Society of Japan 41 (2): 279-88. 
  85. J.B. Dunlop; J.M. Williams; J. Crangle (January-March 1977). "119Sn Mössbauer and neutron diffraction investigation of β Mn-Sn solid solutions". Physica B+C 86-88: 269-71. doi:10.1016/0378-4363(77)90310-2. http://www.sciencedirect.com/science/article/pii/0378436377903102. Retrieved 2015-08-19. 
  86. Z. S. Basinski; J. W. Christian (20 May 1954). "A Pressurized High-Temperature Debye-Scherrer Camera, and Its Use to Determine the Structures and Coefficients of Expansion of γ- and δ-manganese". The Royal Society Proceedings A 223 (1155): 554. doi:10.1098/rspa.1954.0136. http://rspa.royalsocietypublishing.org/content/223/1155/554.short. Retrieved 2015-08-19. 
  87. Griffith, W. P. (2008). "The Periodic Table and the Platinum Group Metals". Platinum Metals Review 52 (2): 114–119. doi:10.1595/147106708X297486. 
  88. 88.0 88.1 Silva, Robert J. (2006). "Fermium, Mendelevium, Nobelium, and Lawrencium" (PDF). In Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and Transactinide Elements. Vol. 3 (3rd ed.). Dordrecht: Springer. pp. 1621–1651. doi:10.1007/1-4020-3598-5_13. {{ISBN|978-1-4020-3555-5 ]]. Archived from the original (PDF) on 2010-07-17.
  89. 89.0 89.1 89.2 89.3 Audi, Georges; Bersillon, Olivier; Blachot, Jean; Wapstra, Aaldert Hendrik (2003), "The NUBASE evaluation of nuclear and decay properties", Nuclear Physics A, 729: 3–128, Bibcode:2003NuPhA.729....3A, doi:10.1016/j.nuclphysa.2003.11.001
  90. 90.0 90.1 Nucleonica (2007–2014). "Universal Nuclide Chart". Nucleonica. Retrieved 22 May 2011.
  91. Eichler, Robert (2013). "First foot prints of chemistry on the shore of the Island of Superheavy Elements". Journal of Physics: Conference Series (IOP Science) 420 (1): 012003. doi:10.1088/1742-6596/420/1/012003. 
  92. "Molybdenum". AZoM.com Pty. Limited. 2007. Retrieved 2007-05-06. {{cite web}}: |archive-date= requires |archive-url= (help)
  93. Toshiba Develops Dysprosium-free Samarium-Cobalt Magnet to Replace Heat-resistant Neodymium Magnet in Essential Applications. Toshiba (2012-08-16). Retrieved on 2012-09-24.
  94. Stamenov P. (2021) Magnetism of the Elements. In: Coey J.M.D., Parkin S.S. (eds) Handbook of Magnetism and Magnetic Materials. Springer, Cham. https://doi.org/10.1007/978-3-030-63210-6_15
  95. Zhang, W., Liu, G. & Han, K. The Fe-Nd (Iron-Neodymium) system. JPE 13, 645–648 (1992). https://doi.org/10.1007/BF02667216
  96. Bala, H., Szymura, S., Pawłowska, G. et al. Effect of impurities on the corrosion behaviour of neodymium. J Appl Electrochem 23, 1017–1024 (1993). https://doi.org/10.1007/BF00266123
  97. Theodore Gray (20 September 2005). "An example of the element Neptunium". Periodic Table.com. Retrieved 2015-11-05.
  98. Yoshida, Zenko; Johnson, Stephen G.; Kimura, Takaumi; Krsul, John R. (2006). Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean. eds. The Chemistry of the Actinide and Transactinide Elements. 3 (3rd ed.). Dordrecht, the Netherlands: Springer. pp. 699–812. doi:10.1007/1-4020-3598-5_6. https://web.archive.org/web/20180117190715/http://radchem.nevada.edu/classes/rdch710/files/neptunium.pdf. 
  99. 99.0 99.1 99.2 P. Swings (July 1943). "Edlén's Identification of the Coronal Lines with Forbidden Lines of Fe X, XI, XIII, XIV, XV; Ni XII, XIII, XV, XVI; Ca XII, XIII, XV; a X, XIV". The Astrophysical Journal 98 (07): 116-28. doi:10.1086/144550. 
  100. D.G. Rancourt and R.B. Scorzelli. Low Spin γ-Fe-Ni (γLS) Proposed as a New Mineral in Fe-Ni-Bearing Meteorites: Epitaxial Intergrowth of γLS and Tetrataenite as Possible Equilibrium State at ~20-40 at % Ni. Journal of Magnetism and Magnetic Materials 150 (1995) 30-36
  101. Beiser, A. (2003). Concepts of modern physics (6th ed.). McGraw-Hill. ISBN 978-0-07-244848-1. OCLC 48965418.
  102. Staszczak, A.; Baran, A.; Nazarewicz, W. (2013). "Spontaneous fission modes and lifetimes of superheavy elements in the nuclear density functional theory". Physical Review C 87 (2): 024320–1. doi:10.1103/physrevc.87.024320. ISSN 0556-2813. 
  103. 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.
  104. Beiser, A. (2003). Concepts of modern physics (6th ed.). McGraw-Hill. ISBN 978-0-07-244848-1. OCLC 48965418.
  105. Oganessian, Yu. Ts.; Rykaczewski, K. P. (2015). "A beachhead on the island of stability". Physics Today 68 (8): 32–38. doi:10.1063/PT.3.2880. ISSN 0031-9228. https://www.osti.gov/biblio/1337838. 
  106. Grant, A. (2018). "Weighing the heaviest elements". Physics Today. doi:10.1063/PT.6.1.20181113a. 
  107. Howes, L. (2019). "Exploring the superheavy elements at the end of the periodic table". Chemical & Engineering News. Retrieved 2020-01-27.
  108. 108.0 108.1 Robinson, A. E. (2019). "The Transfermium Wars: Scientific Brawling and Name-Calling during the Cold War". Distillations. https://www.sciencehistory.org/distillations/the-transfermium-wars-scientific-brawling-and-name-calling-during-the-cold-war. Retrieved 2020-02-22. 
  109. "Популярная библиотека химических элементов. Сиборгий (экавольфрам)" [Popular library of chemical elements. Seaborgium (eka-tungsten)]. n-t.ru (in Russian). Retrieved 2020-01-07. Reprinted from "Экавольфрам". Популярная библиотека химических элементов. Серебро — Нильсборий и далее (in ru). Nauka. 1977. 
  110. Hyde, E. K.; Hoffman, D. C.; Keller, O. L. (1987). "A History and Analysis of the Discovery of Elements 104 and 105". Radiochimica Acta 42 (2): 67–68. doi:10.1524/ract.1987.42.2.57. ISSN 2193-3405. http://www.escholarship.org/uc/item/05x8w9h7. 
  111. Stysziński, Jacek (2010). "Why do we Need Relativistic Computational Methods?". Relativistic Methods for Chemists. Challenges and Advances in Computational Chemistry and Physics. 10. pp. 139–146. doi:10.1007/978-1-4020-9975-5_3. ISBN 978-1-4020-9974-8. 
  112. 112.0 112.1 Hoffman, Darleane C.; Lee, Diana M.; Pershina, Valeria (2006). "Transactinides and the future elements". In Morss; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Dordrecht, The Netherlands: Springer Science+Business Media. ISBN 978-1-4020-3555-5.
  113. Han, Young-Kyu; Bae, Cheolbeom; Son, Sang-Kil; Lee, Yoon Sup (2000). "Spin–orbit effects on the transactinide p-block element monohydrides MH (M=element 113–118)". Journal of Chemical Physics 112 (6): 2684. doi:10.1063/1.480842. https://semanticscholar.org/paper/bb2beba2bc47c3381ed69b7125f6fbfb4b98596a. 
  114. Seth, Michael; Schwerdtfeger, Peter; Fægri, Knut (1999). "The chemistry of superheavy elements. III. Theoretical studies on element 113 compounds". Journal of Chemical Physics 111 (14): 6422–6433. doi:10.1063/1.480168. https://semanticscholar.org/paper/9aaa02788fada5fa2fe16de76ad51718d31b68f8. 
  115. Fricke, Burkhard (1975). "Superheavy elements: a prediction of their chemical and physical properties". Recent Impact of Physics on Inorganic Chemistry. Structure and Bonding. 21: 89–144. doi:10.1007/BFb0116498. ISBN 978-3-540-07109-9. Retrieved 4 October 2013.
  116. Tither, Geoffrey (2001). Minerals, Metals and Materials Society. ed. Progress in Niobium Markets and Technology 1981–2001. https://web.archive.org/web/20081217100553/http://www.cbmm.com.br/portug/sources/techlib/science_techno/table_content/images/pdfs/oppening.pdf. 
  117. Dufresne, Claude; Goyette, Ghislain (2001). Minerals, Metals and Materials Society. ed. The Production of Ferroniobium at the Niobec mine 1981–2001. https://web.archive.org/web/20081217100559/http://www.cbmm.com.br/portug/sources/techlib/science_techno/table_content/sub_1/images/pdfs/start.pdf. 
  118. "ASTM A572 / A572M-18, Standard Specification for High-Strength Low-Alloy Columbium-Vanadium Structural Steel". ASTM International, West Conshohocken. 2018. Retrieved 2020-02-12.
  119. Heisterkamp, Friedrich; Carneiro, Tadeu (2001). Minerals, Metals and Materials Society. ed. Niobium: Future Possibilities – Technology and the Market Place. https://web.archive.org/web/20081217100604/http://www.cbmm.com.br/portug/sources/techlib/science_techno/table_content/images/pdfs/closing.pdf. 
  120. 120.0 120.1 120.2 "Carlsbergite". Webmineral. Retrieved 10 January 2013.
  121. Carlsbergite in the Handbook of Mineralogy
  122. 122.0 122.1 122.2 122.3 Silva, Robert J. (2011). "Chapter 13. Fermium, Mendelevium, Nobelium, and Lawrencium". In Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and Transactinide Elements. Netherlands: Springer. pp. 1621–1651. doi:10.1007/978-94-007-0211-0_13. ISBN 978-94-007-0210-3.
  123. Martin, William C.; Hagan, Lucy; Reader, Joseph; Sugar, Jack (1974). "Ground Levels and Ionization Potentials for Lanthanide and Actinide Atoms and Ions". Journal of Physical and Chemical Reference Data 3 (3): 771–9. doi:10.1063/1.3253147. https://web.archive.org/web/20200215124722/https://pdfs.semanticscholar.org/9618/febdd51cee0e84ff7af88767be47cfcd4818.pdf. 
  124. Lide, David R. (editor), CRC Handbook of Chemistry and Physics, 84th Edition, CRC Press, Boca Raton (FL), 2003, section 10, Atomic, Molecular, and Optical Physics; Ionization Potentials of Atoms and Atomic Ions
  125. 125.0 125.1 Nash, Clinton S. (2005). "Atomic and Molecular Properties of Elements 112, 114, and 118". Journal of Physical Chemistry A. 109 (15): 3493–3500. Bibcode:2005JPCA..109.3493N. doi:10.1021/jp050736o. PMID 16833687.
  126. Haynes, William M., ed. (2011). CRC Handbook of Chemistry and Physics (92nd ed.). CRC Press. ISBN 978-1439855119.
  127. 127.0 127.1 Hudson Institute of Mineralogy (29 October 2015). Palladium. Mindat. http://www.mindat.org/min-3067.html. Retrieved 2015-11-04. 
  128. Verryn, Sabine M. C.; Merkle, Roland K. W. (1994). "Compositional variation of cooperite, braggite, and vysotskite from the Bushveld Complex". Mineralogical Magazine 58 (2): 223–234. doi:10.1180/minmag.1994.058.391.05. 
  129. Genkin, A. D.; Evstigneeva, T. L. (1986). "Associations of platinum- group minerals of the Norilsk copper-nickel sulfide ores". Economic Geology 81 (5): 1203–1212. doi:10.2113/gsecongeo.81.5.1203. 
  130. "Mindat.org - Mines, Minerals and More". www.mindat.org.
  131. 131.0 131.1 131.2 A. Holleman; N. Wiberg (1985). "XV 2.1.3". Lehrbuch der Anorganischen Chemie (33rd ed.). de Gruyter. 
  132. Abundance. ptable.com
  133. Berger, L. I. (1996). Semiconductor materials. CRC Press. p. 84]. ISBN 0-8493-8912-7. https://archive.org/details/semiconductormat0000berg. 
  134. "Allotropes of phosphorus". San Francisco, California: Wikimedia Foundation, Inc. 20 March 2013. Retrieved 2013-03-20.
  135. Roland W. Scholz, ed (2014-03-12). Sustainable Phosphorus Management: A Global Transdisciplinary Roadmap. Springer Science & Business Media. p. 175. ISBN 978-9400772502. 
  136. Mel Schwartz (2016-07-06). Encyclopedia and Handbook of Materials, Parts and Finishes. CRC Press. ISBN 978-1138032064. 
  137. Joseph R. Davisz, ed (January 2001). Copper and Copper Alloys. ASM International. p. 181. 
  138. 138.0 138.1 138.2 T. A. Gornostaeva; P. M. Kartashov; A. V. Mokhov; O. A. Bogatikov (2012). "Native Rhodium-Bearing Ferroplatinum in a Lunar Regolith Sample from the Mare Fecunditatis". Doklady Earth Sciences 444 (2): 770-2. doi:10.1134/S1028334X12060220. http://link.springer.com/article/10.1134/S1028334X12060220#/page-1. Retrieved 2015-11-04. 
  139. Theodore W. Gray (19 October 2002). "Natural plutonium-containing mineral". Periodic Table.com. Retrieved 2015-11-05.
  140. Hecker, Siegfried S. (2000). "Plutonium and its alloys: from atoms to microstructure". Los Alamos Science 26: 290–335. https://web.archive.org/web/20090224204042/http://www.fas.org/sgp/othergov/doe/lanl/pubs/00818035.pdf. Retrieved February 15, 2009. 
  141. CST (20 November 2000). "The Simple Cubic Lattice". Washington, DC USA: The Naval Research Laboratory. Retrieved 2015-08-27.
  142. CSTPo (20 November 2000). "The A_i (beta Po) Structure". Washington, DC USA: The Naval Research Laboratory. Retrieved 2015-08-27.
  143. 143.0 143.1 Lorence G. Collins (3 February 1997). "Polonium Halos and Myrmekite in Pegmatite and Granite" (PDF). Northridge, California USA: California State University, Northridge. Retrieved 2015-08-27.
  144. 144.0 144.1 P. G. Nutting (January 1906). "Line Structure. I.". The Astrophysical Journal 23 (1): 64-78. doi:10.1086/141302. 
  145. Rhodes, Richard (2002). Dark Sun: The Making of the Hydrogen Bomb. New York: Walker & Company. pp. 187–188]. https://archive.org/details/darksunmakingofh00rhod/page/187. 
  146. Weigel, F. (1959). "Chemie des Poloniums". Angewandte Chemie 71 (9): 289–316. doi:10.1002/ange.19590710902. 
  147. Pallmer, P. G.; Chikalla, T. D. (1971). "The crystal structure of promethium". Journal of the Less Common Metals 24 (3): 233. doi:10.1016/0022-5088(71)90101-9. 
  148. 148.0 148.1 Gschneidner Jr., K.A. (2005). Lide, D. R.. ed. Physical Properties of the rare earth metals, In: CRC Handbook of Chemistry and Physics (86th ed.). Boca Raton, FL: CRC Press. https://link.springer.com/article/10.1007/BF03029283. Retrieved 2012-06-20. 
  149. "Protactinium, Human Health Fact Sheet" (PDF). Argonne National Laboratory. August 2005. Retrieved 7 March 2008.
  150. 150.0 150.1 David A. Young (11 September 1975). Phase Diagrams of the Elements. University of California, Livermore, California USA: Lawrence Livermore Laboratory. pp. 70. http://www.iaea.org/inis/collection/NCLCollectionStore/_Public/07/255/7255152.pdf. Retrieved 2015-08-26. 
  151. A. F. Grachev; S. E. Borisovsky; A. V. Grigor’eva (October 2008). "The first find of native rhenium in the transitional clay layer at the Cretaceous/Paleogene boundary in the Gams Section (eastern Alps, Austria)". Doklady Earth Sciences 422 (1): 1065-7. doi:10.1134/S1028334X08070131. http://link.springer.com/article/10.1134/S1028334X08070131. Retrieved 2015-11-04. 
  152. A. R. Hildebrand; W. V. Boynton; W. H. Zoller (June 1984). "Rhenium Enriched Kilauea Volcano Aerosols: Evidence for a Volcanogenic Component in the K/T Boundary Clay Layer". Bulletin of the American Astronomical Society 16 (06): 679. http://adsabs.harvard.edu/abs/1984BAAS...16..679H. Retrieved 2014-01-10. 
  153. Cramer, Stephen D.; Covino, Jr., Bernard S., eds (1990). ASM handbook. Materials Park, OH: ASM International. pp. 393–396. https://books.google.com/books?id=QV0sWU2qF5oC&pg=PA396. 
  154. A. Marinov, A. Pape, D. Kolb, L. Halicz, I. Segal, N. Tepliakov and R. Brandt (2011). "Enrichment of the Superheavy Element Roentgenium (Rg) in Natural Au". International Journal of Modern Physics E 20 (11): 2391-2401. doi:10.1142/S0218301311020393. http://www.phys.huji.ac.il/~marinov/publications/Rg_261_arXiv_77.pdf. Retrieved 2014-04-08. 
  155. Aksenov, Nikolay V.; Steinegger, Patrick; Abdullin, Farid Sh.; Albin, Yury V.; Bozhikov, Gospodin A.; Chepigin, Viktor I.; Eichler, Robert; Lebedev, Vyacheslav Ya. et al. (July 2017). "On the volatility of nihonium (Nh, Z = 113)". The European Physical Journal A 53 (158): 158. doi:10.1140/epja/i2017-12348-8. https://www.semanticscholar.org/paper/5a07c41dfd0fc2913510dc843a5bc9a506bc92d4. 
  156. Haynes, William M., ed. (2016). CRC Handbook of Chemistry and Physics (97th ed.). CRC Press. ISBN 9781498754293.
  157. Hartman, H. L., ed (1992). SME mining engineering handbook. Littleton, Colo.: Society for Mining, Metallurgy, and Exploration. p. 69. https://books.google.com/books?id=Wm6QMRaX9C4C&pg=PA69. 
  158. Harris, Donald C.; Cabri, Louis J. (1 August 1973). "The nomenclature of the natural alloys of osmium, iridium and ruthenium based on new compositional data of alloys from world-wide occurrences". The Canadian Mineralogist 12 (2): 104–112. https://pubs.geoscienceworld.org/canmin/article-abstract/12/2/104/10913/The-nomenclature-of-the-natural-alloys-of-osmium. 
  159. Ghiorso, A.; Nurmia, M.; Harris, J.; Eskola, K.; Eskola, P. (1969). "Positive Identification of Two Alpha-Particle-Emitting Isotopes of Element 104". Physical Review Letters 22 (24): 1317–1320. doi:10.1103/PhysRevLett.22.1317. https://cloudfront.escholarship.org/dist/prd/content/qt3fm666nq/qt3fm666nq.pdf. 
  160. Bemis, C. E.; Silva, R.; Hensley, D.; Keller, O.; Tarrant, J.; Hunt, L.; Dittner, P.; Hahn, R. et al. (1973). "X-Ray Identification of Element 104". Physical Review Letters 31 (10): 647–650. doi:10.1103/PhysRevLett.31.647. 
  161. 161.0 161.1 Shi, N.; Fort, D. (1985). "Preparation of samarium in the double hexagonal close packed form". Journal of the Less Common Metals 113 (2): 21. doi:10.1016/0022-5088(85)90294-2. 
  162. Vohra, Y.; Akella, Jagannadham; Weir, Sam; Smith, Gordon S. (1991). "A new ultra-high pressure phase in samarium". Physics Letters A 158 (1–2): 89. doi:10.1016/0375-9601(91)90346-A. https://zenodo.org/record/1258493. 
  163. Leger, J.; Yacoubi, N.; Loriers, J. (1981). "Synthesis of rare earth monoxides". Journal of Solid State Chemistry 36 (3): 261. doi:10.1016/0022-4596(81)90436-9. 
  164. Brown, R.; Clark, N. J. (1974). "Composition limits and vaporization behaviour of rare earth nitrides". Journal of Inorganic and Nuclear Chemistry 36 (11): 2507. doi:10.1016/0022-1902(74)80462-8. 
  165. Meng, J.; Ren, Yufang (1991). "Studies on the electrical properties of rare earth monophosphides". Journal of Solid State Chemistry 95 (2): 346. doi:10.1016/0022-4596(91)90115-X. 
  166. Beeken, R.; Schweitzer, J. (1981). "Intermediate valence in alloys of SmSe with SmAs". Physical Review B 23 (8): 3620. doi:10.1103/PhysRevB.23.3620. 
  167. Burrell, A. Willey Lower "Aluminum scandium alloy" U.S. Patent 3,619,181 issued on November 9, 1971.
  168. Zakharov, V. V. (2003). "Effect of Scandium on the Structure and Properties of Aluminum Alloys". Metal Science and Heat Treatment 45 (7/8): 246. doi:10.1023/A:1027368032062. 
  169. Youssef, Khaled M.; Zaddach, Alexander J.; Niu, Changning; Irving, Douglas L.; Koch, Carl C. (2015). "A Novel Low-Density, High-Hardness, High-entropy Alloy with Close-packed Single-phase Nanocrystalline Structures". Materials Research Letters 3 (2): 95–99. doi:10.1080/21663831.2014.985855. 
  170. J. R. Kline; J. E. Foss; S. S. Brar (March 1969). "Lanthanum and Scandium Distribution in Three Glacial Soils of Western Wisconsin". Soil Science Society of America Journal 33 (2): 287-91. doi:10.2136/sssaj1969.03615995003300020034x. https://dl.sciencesocieties.org/publications/sssaj/abstracts/33/2/SS0330020287. Retrieved 2014-10-01. 
  171. 171.0 171.1 Sonzogni, Alejandro. "Interactive Chart of Nuclides". National Nuclear Data Center: Brookhaven National Laboratory. Archived from the original on 2018-06-12. Retrieved 2008-06-06.
  172. 172.0 172.1 Gray, Theodore (2002–2010). "The Photographic Periodic Table of the Elements". periodictable.com. Retrieved 16 November 2012.
  173. "β –Se (Al ) Structure: A_mP32_14_8e". Encyclopedia of Crystallographic Prototypes.
  174. "β –Se (Al ) Structure: A_mP32_14_16e". Encyclopedia of Crystallographic Prototypes.
  175. Olav Foss and Vitalijus Janickis (1980). "Crystal structure of γ-monoclinic selenium". Journal of the Chemical Society, Dalton Transactions (4): 624–627. doi:10.1039/DT9800000624. 
  176. House, James E. (2008). Inorganic chemistry. Academic Press. p. 524. ISBN 978-0-12-356786-4. 
  177. 177.0 177.1 Davis, Joseph R. (2001). Copper and Copper Alloys. ASM Int.. p. 91. https://books.google.com/books?id=sxkPJzmkhnUC&pg=PA91. 
  178. Isakov, Edmund (2008-10-31). Cutting Data for Turning of Steel. p. 67. https://books.google.com/books?id=QahG1Ou1cyEC&pg=PA67. 
  179. Gol'Dshtein, Ya. E.; Mushtakova, T. L.; Komissarova, T. A. (1979). "Effect of selenium on the structure and properties of structural steel". Metal Science and Heat Treatment 21 (10): 741–746. doi:10.1007/BF00708374. 
  180. 180.0 180.1 Warr, L.N. (2021). IMA–CNMNC approved mineral symbols. Mineralogical Magazine, 85(3), 291-320. doi:10.1180/mgm.2021.43
  181. Mindat Profile
  182. Achavalite data on WebMineral
  183. Hålenius, U., Hatert, F., Pasero, M., and Mills, S.J., IMA Commission on New Minerals, Nomenclature and Classification (CNMNC) Newsletter 28. Mineralogical Magazine 79(7), 1859–1864
  184. "siderophile". San Francisco, California: Wikimedia Foundation, Inc. 19 June 2013. Retrieved 2015-02-19.
  185. Horton E. Newsom (13 October 1986). W. K. Hartmann. ed. Constraints on the origin of the Moon from the abundance of molybdenum and other siderophile elements, In: Origin of the moon. Kona, HI USA: Lunar and Planetary Institute. pp. 203-29. http://adsabs.harvard.edu/cgi-bin/nph-data_query?bibcode=1986ormo.conf..203N&link_type=ARTICLE&db_key=AST&high=. Retrieved 2016-10-31. 
  186. C.W. Dale, K.W. Burton, D.G. Pearson, A. Gannoun, O. Alard, T.W. Arglesb, and I.J. Parkinson (2009). "The behaviour of highly siderophile elements in oceanic crust during subduction: whole-rock and mineral-scale insights from a high-pressure terrain". Geochimica et Cosmochimica Acta 73 (5): 1394-416. doi:10.1016/j.gca.2008.11.036. http://dro.dur.ac.uk/10680/1/10680.pdf. Retrieved 2016-10-31. 
  187. Eiji Ohtani (10 January 2017). "New candidate for 'missing element' in Earth's core". London, England: BBC. Retrieved 2017-01-11.
  188. 188.0 188.1 Rebecca Morelle (10 January 2017). "New candidate for 'missing element' in Earth's core". London, England: BBC. Retrieved 2017-01-11.
  189. 189.0 189.1 189.2 Simon Redfern (10 January 2017). "New candidate for 'missing element' in Earth's core". London, England: BBC. Retrieved 2017-01-11.
  190. Peter S. Conti; Eva M. Leep (October 1974). "Spectroscopic observations of O-type stars. V. The hydrogen lines and lambda 4686 He II". The Astrophysical Journal 193 (10): 113-24. doi:10.1086/153135. 
  191. Apelian, D. (2009). "Aluminum Cast Alloys: Enabling Tools for Improved Performance" (PDF). Wheeling, Illinois: North American Die Casting Association. {{cite web}}: |archive-date= requires |archive-url= (help)
  192. Corathers, Lisa A. 2009 Minerals Yearbook. USGS
  193. D. Lal; B.L.K. Somayajulu (June 1984). "Some aspects of the geochemistry of silicon isotopes". Tectonophysics 105 (1-4): 383-97. doi:10.1016/0040-1951(84)90215-4. http://www.sciencedirect.com/science/article/pii/0040195184902154. Retrieved 2014-09-30. 
  194. 194.0 194.1 T. L. Smith; A. N. Witt (December 1999). "The Photoluminescence Efficiency of Extended Red Emission as a Constraint for Interstellar Dust". Bulletin of the American Astronomical Society 31: 1479. http://adsabs.harvard.edu/abs/1999AAS...195.7406S. Retrieved 2013-08-02. 
  195. Cameron P. Wake; Paul Andrew Mayewski; Xie Zichu; Wang Ping; Li Zhongquin (July 23 1993). "Regional Distribution of Monsoon and Desert Dust Signals Recorded in Asian Glaciers". Geophysical Research Letters 20 (14): 1411-4. http://digitalcommons.library.umaine.edu/cgi/viewcontent.cgi?article=1190&context=ers_facpub&sei-redir=1&referer=http%3A%2F%2Fscholar.google.com%2Fscholar%3Fq%3Dsodium%2Bglaciers%2B-acetic%2B-apples%26btnG%3D%26hl%3Den%26as_sdt%3D0%252C3#search=%22sodium%20glaciers%20-acetic%20-apples%22. Retrieved 2014-09-29. 
  196. Ropp, Richard C. (31 December 2012). Encyclopedia of the Alkaline Earth Compounds. p. 16. ISBN 978-0-444-59553-9. https://books.google.com/books?id=yZ786vEild0C&pg=PA16. 
  197. Holleman, A. F.; Wiberg, E.; Wiberg, N. (2007). Lehrbuch der Anorganischen Chemie (in de) (102nd ed.). de Gruyter. 
  198. Cowan, Clyde L., Jr.; Reines, Frederick; Harrison, Francis B. "Kiko"; Kruse, Herald W.; McGuire, Austin D. (1956). "Detection of the free neutrino: A confirmation". Science 124 (3212): 103–104. doi:10.1126/science.124.3212.103. PMID 17796274. 
  199. Nozariasbmarz, Amin; Poudel, Bed; Li, Wenjie; Kang, Han Byul; Zhu, Hangtian; Priya, Shashank (2020-07-24). "Bismuth Telluride Thermoelectrics with 8% Module Efficiency for Waste Heat Recovery Application". iScience 23 (7): 101340. doi:10.1016/j.isci.2020.101340. ISSN 2589-0042. https://www.sciencedirect.com/science/article/pii/S2589004220305277. 
  200. Leddicotte, G. W. (1961). The radiochemistry of tellurium. Nuclear science series. Subcommittee on Radiochemistry, National Academy of Sciences-National Research Council. p. 5. http://library.lanl.gov/cgi-bin/getfile?rc000049.pdf. 
  201. Oganessian, Yu.Ts.; Abdullin, F.Sh.; Bailey, P.D.; Benker, D.E.; Bennett, M.E.; Dmitriev, S.N.; Ezold, J.G.; Hamilton, J.H. et al. (2010). "Synthesis of a new element with atomic number Z = 117". Physical Review Letters 104 (14): 142502. doi:10.1103/PhysRevLett.104.142502. PMID 20481935. https://semanticscholar.org/paper/ec9412add23e66f34b6bf51ebd7332278af413fc. 
  202. Patnaik, Pradyot (2003). Handbook of Inorganic Chemical Compounds. McGraw-Hill. pp. 920–921. https://books.google.com/books?id=Xqj-TTzkvTEC&pg=PA243. Retrieved 2009-06-06. 
  203. Hudson Institute of Mineralogy (1993–2018). "Mindat.org". www.mindat.org. Retrieved 14 January 2018.
  204. Wickleder, Mathias S.; Fourest, Blandine; Dorhout, Peter K. (2006). "Thorium". The Chemistry of the Actinide and Transactinide Elements. pp. 52–160. doi:10.1007/1-4020-3598-5_3. ISBN 978-1-4020-3555-5. 
  205. Tretyakov, Yu. D., ed (2007). Non-organic chemistry in three volumes. Chemistry of transition elements. 3. Academy. 
  206. Monazite group on Mindat.org
  207. "Helium From Sand", March 1931, Popular Mechanics p. 460.
  208. "Umbozerite". Retrieved 19 November 2021.
  209. "Umbozerite Mineral Data". Retrieved 19 November 2021.
  210. 210.0 210.1 210.2 "Umbozerite" (PDF). Retrieved 19 November 2021.
  211. "Chemical reactions of Thulium". Webelements. Retrieved 2009-06-06.
  212. Holleman, Arnold F.; Wiberg, Egon; Wiberg, Nils (1985). "Tin". Lehrbuch der Anorganischen Chemie (in de) (91–100 ed.). Walter de Gruyter. pp. 793–800. 
  213. Schwartz, Mel (2002). "Tin and alloys, properties". Encyclopedia of Materials, Parts and Finishes (2nd ed.). CRC Press. 
  214. Molodets, A.M.; Nabatov, S.S. (2000). "Thermodynamic potentials, diagram of state, and phase transitions of tin on shock compression". High Temperature 38 (5): 715–721. doi:10.1007/BF02755923. 
  215. Jing Chen, Jiliang Li, and Jun Wu (30 April 2000). "Native titanium inclusions in the coesite eclogites from Dabieshan, China". Earth and Planetary Science Letters 177 (3-4): 237-40. doi:10.1016/S0012-821X(00)00057-1. http://www.sciencedirect.com/science/article/pii/S0012821X00000571. Retrieved 2015-08-19. 
  216. B.B. Panigrahi, M.M. Godkhindi , K. Das, P.G. Mukunda, and P. Ramakrishnan (15 April 2005). "Sintering kinetics of micrometric titanium powder". Materials Science and Engineering: A 396 (1-2): 255-62. doi:10.1016/j.msea.2005.01.016. http://www.sciencedirect.com/science/article/pii/S0921509305000778. Retrieved 2015-08-19. 
  217. G. Stellmacher; E. Wiehr (August 1991). "Geometric line elevation in solar limb faculae". Astronomy and Astrophysics 248 (1): 227-31. 
  218. G. Catanzaro (January 2010). "First spectroscopic analysis of β Scorpii C and β Scorpii E Discovery of a new HgMn star in the multiple system β Scorpii". Astronomy & Astrophysics 509: 7. doi:10.1051/0004-6361/200913332. http://www.aanda.org/articles/aa/pdf/2010/01/aa13332-09.pdf. Retrieved 2013-01-18. 
  219. "Osbornite". Mindat.org. Hudson Institute of Mineralogy. Retrieved February 29, 2016.
  220. "Osbornite Mineral Data". Mineralogy Database. David Barthelmy. September 5, 2012. Retrieved October 6, 2015.
  221. Li Ziying, Huang Zhizhang, Li Xiuzhen, Guo Jian, Fan Chou (24 October 2015). "The Discovery of Natural Native Uranium and Its Significance". Acta Geologica Sinica 89 (5): 1561-1567. doi:10.1111/1755-6724.12564. https://onlinelibrary.wiley.com/doi/abs/10.1111/1755-6724.12564. Retrieved 29 June 2022. 
  222. "Development of DU Munitions". Depleted Uranium in the Gulf (II). Gulflink, official website of Force Health Protection & Readiness. 2000.
  223. B. S. Chandrasekhar and J. K. Hulm (November 1958). "The electrical resistivity and super-conductivity of some uranium alloys and compounds". Journal of Physics and Chemistry of Solids 7 (2-3): 259-267. doi:10.1016/0022-3697(58)90271-3. https://www.sciencedirect.com/science/article/abs/pii/0022369758902713. Retrieved 29 June 2022. 
  224. Rolf-Dieter Hoffmann, Rainer Pöttgen, Bernard Chevalier, Etienne Gaudin, and Samir F. Matar (July 2013). "The ternary germanides UMnGe and U
    2
    Mn
    3
    Ge
    "
    . Solid State Sciences 21: 73-80. doi:10.1016/j.solidstatesciences.2013.04.006. https://www.sciencedirect.com/science/article/abs/pii/S1293255813001301. Retrieved 29 June 2022.
     
  225. 225.0 225.1 MikhailI Ostrooumov and Yuri Taran (20 May 2015). "Discovery of Native Vanadium, a New Mineral from the Colima Volcano, State of Colima (Mexico)". Revista de la Sociedad Española de Mineralogía: 109-10. http://www.uhu.es/fexp/sem2015/arc/macla/macla_20_109-110.pdf. Retrieved 2015-08-19. 
  226. 226.0 226.1 Bucher, E.; Schmidt, P.; Jayaraman, A.; Andres, K.; Maita, J.; Nassau, K.; Dernier, P. (1970). "New First-Order Phase Transition in High-Purity Ytterbium Metal". Physical Review B 2 (10): 3911. doi:10.1103/PhysRevB.2.3911. 
  227. Jackson, M. (2000). "Magnetism of Rare Earth". The IRM quarterly 10(3): 1
  228. 228.0 228.1 "Nucleonica: Universal Nuclide Chart". Nucleonica. 2007–2011. Retrieved July 22, 2011.
  229. G. H. Herbig (March 1974). "VY Canis Majoris. IV. The emission bands of ScO". The Astrophysical Journal 188 (3): 533-8. doi:10.1086/152744. 
  230. "Yttrium". Periodic Table of Elements: LANL. Los Alamos National Security. http://periodic.lanl.gov/39.shtml. 
  231. 231.0 231.1 231.2 Lide, David R., ed (2007–2008). "Yttrium". CRC Handbook of Chemistry and Physics. 4. New York: CRC Press. p. 41. 
  232. 232.0 232.1 Daane, A. H. (1968). "Yttrium". In Hampel, Clifford A. (ed.). The Encyclopedia of the Chemical Elements. New York: Reinhold Book Corporation. pp. 810–821. LCCN 68029938. OCLC 449569.
  233. Ingalls, Walter Renton (1902). Production and Properties of Zinc: A Treatise on the Occurrence and Distribution of Zinc Ore, the Commercial and Technical Conditions Affecting the Production of the Spelter, Its Chemical and Physical Properties and Uses in the Arts, Together with a Historical and Statistical Review of the Industry. The Engineering and Mining Journal. pp. 142–6. https://books.google.com/books?id=RhNDAAAAIAAJ&pg=PA133. 
  234. David R. Lide, ed. (2006). Handbook of Chemistry and Physics (87th ed.). Boca Raton, Florida: CRC Press, Taylor & Francis Group. ISBN 978-0-8493-0487-3.
  235. 235.0 235.1 Craddock, P.T. and Eckstein, K (2003) "Production of Brass in Antiquity by Direct Reduction" in Craddock, P.T. and Lang, J. (eds) Mining and Metal Production Through the Ages London: British Museum pp. 226–7
  236. https://www.mindat.org/min-6830.html
  237. 237.0 237.1 237.2 237.3 http://www.mindat.org/min-4410.html
  238. F. Vigué; E. Tournié; J.-P. Faurie (January 2000). "Zn(Mg)BeSe-based p-i-n photodiodes operating in the blue-violet and near-ultraviolet spectral range". Applied Physics Letters 76 (2): 242-4. doi:10.1063/1.125715. http://ieeexplore.ieee.org/xpls/abs_all.jsp?arnumber=4915124. Retrieved 2013-06-01. 
  239. "Zirconium, In: How Products Are Made". Advameg Inc. 2007. Retrieved 2008-03-26.
  240. Glenn D. Considine, ed (2005). Zirconium, In: Van Nostrand's Encyclopedia of Chemistry. New York: Wylie-Interscience. pp. 1778–1779. ISBN 978-0-471-61525-5. 
  241. David R. Lide, ed (2007). Zirconium, In: CRC Handbook of Chemistry and Physics. 4. New York: CRC Press. p. 42. 
  242. Mark Winter (2007). "Electronegativity (Pauling)". University of Sheffield. Retrieved 2008-03-05.
  243. Schnell I; Albers RC (January 2006). "Zirconium under pressure: phase transitions and thermodynamics". Journal of Physics: Condensed Matter 18 (5): 16. doi:10.1088/0953-8984/18/5/001. 
  244. David L. Lambert; Verne V. Smith; Maurizio Busso; Roberto Gallino; Oscar Straniero (September 1, 1995). "The Chemical Composition of Red Giants. IV. The Neutron Density at the s-Process Site". The Astrophysical Journal 450 (09): 302-17. doi:10.1086/176141. http://adsabs.harvard.edu/abs/1995ApJ...450..302L. Retrieved 2013-08-01. 
  245. 245.0 245.1 245.2 Audi, G; Bersillon, O.; Blachot, J.; Wapstra, A. H. (2003). "Nubase2003 Evaluation of Nuclear and Decay Properties". Nuclear Physics A 729 (1): 3–128. doi:10.1016/j.nuclphysa.2003.11.001. http://hal.in2p3.fr/in2p3-00014184. 
  246. 246.0 246.1 246.2 Emily Conover (January 7, 2019). A weird type of zirconium soaks up neutrons like a sponge. Science News. https://www.sciencenews.org/article/weird-type-zirconium-soaks-neutrons-sponge?utm_source=email&utm_medium=email&utm_campaign=latest-newsletter-v2. Retrieved 9 January 2019. 
  247. Jennifer A. Shusterman; Nicholas D. Scielzo; Keenan J. Thomas; Eric B. Norman; Suzanne E. Lapi; C. Shaun Loveless Nickie J. Peters, J. David Robertson, Dawn A. Shaughnessy & Anton P. Tonchev (7 January 2019). "The surprisingly large neutron capture cross-section of 88Zr". Nature 18 (838): 10. doi:10.1038/s41586-018-0838-z. https://www.nature.com/articles/s41586-018-0838-z.epdf?referrer_access_token=dHufYV4FNXVsGEZyZ6g46dRgN0jAjWel9jnR3ZoTv0Mc5ChkkjTroRA03db24PlRo5oLvqS_cQqr3Ff-LqPf8ZW9FUa-GytgPnedZFlJA3sJHCPQ-Wb0LQHgUooFhbXFT2u0gCjThP3WjpB7tVio9JlGs_SyStggLGcHQu7l_hT2N2TYQtXDBqInN55QTO7c3rRCffeqfGp9nOYIDqm3LtP2DriqXGv9cFWQjji220S1F_HyMnoVXS8PYQKna8D8JsxmnGI5DENoTSqjSgCoRCc8rj1Fc9iU2JqJ1KGsMmzESD_APHFsnyOGfnXNr1URBD6edhDX5R8Ax1g0rQo0mTNUjtaJyadTVh12QnoMqcg%3D&tracking_referrer=www.sciencenews.org. Retrieved 9 January 2019. 
[edit | edit source]

{{Radiation astronomy resources}}{{Repellor vehicle}}