Dysprosium

Last updated
Dysprosium, 66Dy
Dy chips.jpg
Dysprosium
Pronunciation /dɪsˈprziəm/ (dis-PROH-zee-əm)
Appearancesilvery white
Standard atomic weight Ar°(Dy)
Dysprosium in the periodic table
Hydrogen Helium
Lithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine Neon
Sodium Magnesium Aluminium Silicon Phosphorus Sulfur Chlorine Argon
Potassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine Xenon
Caesium Barium Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury (element) Thallium Lead Bismuth Polonium Astatine Radon
Francium Radium Actinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium Rutherfordium Dubnium Seaborgium Bohrium Hassium Meitnerium Darmstadtium Roentgenium Copernicium Nihonium Flerovium Moscovium Livermorium Tennessine Oganesson


Dy

Cf
terbiumdysprosiumholmium
Atomic number (Z)66
Group f-block groups (no number)
Period period 6
Block   f-block
Electron configuration [ Xe ] 4f10 6s2
Electrons per shell2, 8, 18, 28, 8, 2
Physical properties
Phase at  STP solid
Melting point 1680  K (1407 °C,2565 °F)
Boiling point 2840 K(2562 °C,4653 °F)
Density (at 20° C)8.550 g/cm3 [3]
when liquid (at  m.p.)8.37 g/cm3
Heat of fusion 11.06  kJ/mol
Heat of vaporization 280 kJ/mol
Molar heat capacity 27.7 J/(mol·K)
Vapor pressure
P (Pa)1101001 k10 k100 k
at T (K)13781523(1704)(1954)(2304)(2831)
Atomic properties
Oxidation states 0, [4] +1, +2, +3, +4 (a weakly basic oxide)
Electronegativity Pauling scale: 1.22
Ionization energies
  • 1st: 573.0 kJ/mol
  • 2nd: 1130 kJ/mol
  • 3rd: 2200 kJ/mol
Atomic radius empirical:178  pm
Covalent radius 192±7 pm
Dysprosium spectrum visible.png
Spectral lines of dysprosium
Other properties
Natural occurrence primordial
Crystal structure hexagonal close-packed (hcp)(hP2)
Lattice constants
Hexagonal close packed.svg
a = 359.16 pm
c = 565.01 pm (at 20 °C) [3]
Thermal expansion poly: 9.9 µm/(m⋅K)(r.t.)
Thermal conductivity 10.7 W/(m⋅K)
Electrical resistivity α, poly: 926 nΩ⋅m(r.t.)
Magnetic ordering paramagnetic at 300 K
Molar magnetic susceptibility +103500×10−6 cm3/mol(293.2 K) [5]
Young's modulus 61.4 GPa
Shear modulus 24.7 GPa
Bulk modulus 40.5 GPa
Speed of sound thin rod2710 m/s(at 20 °C)
Poisson ratio 0.247
Vickers hardness 410–550 MPa
Brinell hardness 500–1050 MPa
CAS Number 7429-91-6
History
Discovery Lecoq de Boisbaudran (1886)
First isolation Georges Urbain (1905)
Isotopes of dysprosium
Main isotopes [6] Decay
abun­dance half-life (t1/2) mode pro­duct
154Dy synth 1.40×106 y [7] α 150Gd
156Dy0.056% stable
158Dy0.095%stable
160Dy2.33%stable
161Dy18.9%stable
162Dy25.5%stable
163Dy24.9%stable
164Dy28.3%stable
165Dysynth2.334 hβ165Ho
Symbol category class.svg  Category: Dysprosium
| references

Dysprosium is a chemical element; it has symbol Dy and atomic number 66. It is a rare-earth element in the lanthanide series with a metallic silver luster. Dysprosium is never found in nature as a free element, though, like other lanthanides, it is found in various minerals, such as xenotime. Naturally occurring dysprosium is composed of seven isotopes, the most abundant of which is 164Dy.

Contents

Dysprosium was first identified in 1886 by Paul Émile Lecoq de Boisbaudran, but it was not isolated in pure form until the development of ion-exchange techniques in the 1950s. Dysprosium has relatively few applications where it cannot be replaced by other chemical elements. It is used for its high thermal neutron absorption cross-section in making control rods in nuclear reactors, for its high magnetic susceptibility (χv5.44×10−3) in data-storage applications, and as a component of Terfenol-D (a magnetostrictive material). Soluble dysprosium salts are mildly toxic, while the insoluble salts are considered non-toxic.

Characteristics

Physical properties

Dysprosium sample Dysprosium.jpg
Dysprosium sample

Dysprosium is a rare-earth element and has a metallic, bright silver luster. It is quite soft and can be machined without sparking if overheating is avoided. Dysprosium's physical characteristics can be greatly affected by even small amounts of impurities. [8]

Dysprosium and holmium have the highest magnetic strengths of the elements, [9] especially at low temperatures. [10] Dysprosium has a simple ferromagnetic ordering at temperatures below its Curie temperature of 90.5 K (−182.7 °C), at which point it undergoes a first-order phase transition from the orthorhombic crystal structure to hexagonal close-packed (hcp). [3] It then has a helical antiferromagnetic state, in which all of the atomic magnetic moments in a particular basal plane layer are parallel and oriented at a fixed angle to the moments of adjacent layers. This unusual antiferromagnetism transforms into a disordered (paramagnetic) state at 179 K (−94 °C). [11] It transforms from the hcp phase to the body-centered cubic phase at 1,654 K (1,381 °C). [3]

Chemical properties

Dysprosium metal retains its luster in dry air but it will tarnish slowly in moist air, and it burns readily to form dysprosium(III) oxide:

4 Dy + 3 O2 → 2 Dy2O3

Dysprosium is quite electropositive and reacts slowly with cold water (and quickly with hot water) to form dysprosium hydroxide:

2 Dy (s) + 6 H2O (l) → 2 Dy(OH)3 (aq) + 3 H2 (g)

Dysprosium hydroxide decomposes to form DyO(OH) at elevated temperatures, which then decomposes again to dysprosium(III) oxide. [12]

Dysprosium metal vigorously reacts with all the halogens at above 200 °C:[ citation needed ]

2 Dy (s) + 3 F2 (g) → 2 DyF3 (s) [green]
2 Dy (s) + 3 Cl2 (g) → 2 DyCl3 (s) [white]
2 Dy (s) + 3 Br2 (l) → 2 DyBr3 (s) [white]
2 Dy (s) + 3 I2 (g) → 2 DyI3 (s) [green]

Dysprosium dissolves readily in dilute sulfuric acid to form solutions containing the yellow Dy(III) ions, which exist as a [Dy(OH2)9]3+ complex: [13]

2 Dy (s) + 3 H2SO4 (aq) → 2 Dy3+ (aq) + 3 SO2−
4
(aq) + 3 H2 (g)

The resulting compound, dysprosium(III) sulfate, is noticeably paramagnetic.

Compounds

Dysprosium sulfate, Dy2(SO4)3 Dysprosium-sulfate.jpg
Dysprosium sulfate, Dy2(SO4)3

Dysprosium halides, such as DyF3 and DyBr3, tend to take on a yellow color. Dysprosium oxide, also known as dysprosia, is a white powder that is highly magnetic, more so than iron oxide. [10]

Dysprosium combines with various non-metals at high temperatures to form binary compounds with varying composition and oxidation states +3 and sometimes +2, such as DyN, DyP, DyH2 and DyH3; DyS, DyS2, Dy2S3 and Dy5S7; DyB2, DyB4, DyB6 and DyB12, as well as Dy3C and Dy2C3. [14]

Dysprosium carbonate, Dy2(CO3)3, and dysprosium sulfate, Dy2(SO4)3, result from similar reactions. [15] Most dysprosium compounds are soluble in water, though dysprosium carbonate tetrahydrate (Dy2(CO3)3·4H2O) and dysprosium oxalate decahydrate (Dy2(C2O4)3·10H2O) are both insoluble in water. [16] [17] Two of the most abundant dysprosium carbonates, Dy2(CO3)3·2–3H2O (similar to the mineral tengerite-(Y)), and DyCO3(OH) (similar to minerals kozoite-(La) and kozoite-(Nd), are known to form via a poorly ordered (amorphous) precursor phase with a formula of Dy2(CO3)3·4H2O. This amorphous precursor consists of highly hydrated spherical nanoparticles of 10–20 nm diameter that are exceptionally stable under dry treatment at ambient and high temperatures. [18]

Isotopes

Naturally occurring dysprosium is composed of seven isotopes: 156Dy, 158Dy, 160Dy, 161Dy, 162Dy, 163Dy, and 164Dy. These are all considered stable, although only the last two are theoretically stable: the others can theoretically undergo alpha decay. Of the naturally occurring isotopes, 164Dy is the most abundant at 28%, followed by 162Dy at 26%. The least abundant is 156Dy at 0.06%. [19] Dysprosium is the heaviest element to have isotopes that are predicted to be stable rather than observationally stable isotopes that are predicted to be radioactive.

Twenty-nine radioisotopes have been synthesized, ranging in atomic mass from 138 to 173. The most stable of these is 154Dy, with a half-life of approximately 3×106 years, followed by 159Dy with a half-life of 144.4 days. The least stable is 138Dy, with a half-life of 200 ms. As a general rule, isotopes that are lighter than the stable isotopes tend to decay primarily by β+ decay, while those that are heavier tend to decay by β decay. However, 154Dy decays primarily by alpha decay, and 152Dy and 159Dy decay primarily by electron capture. [19] Dysprosium also has at least 11 metastable isomers, ranging in atomic mass from 140 to 165. The most stable of these is 165mDy, which has a half-life of 1.257 minutes. 149Dy has two metastable isomers, the second of which, 149m2Dy, has a half-life of 28 ns. [19]

History

In 1878, erbium ores were found to contain the oxides of holmium and thulium. French chemist Paul Émile Lecoq de Boisbaudran, while working with holmium oxide, separated dysprosium oxide from it in Paris in 1886. [20] [21] His procedure for isolating the dysprosium involved dissolving dysprosium oxide in acid, then adding ammonia to precipitate the hydroxide. He was only able to isolate dysprosium from its oxide after more than 30 attempts at his procedure. On succeeding, he named the element dysprosium from the Greek dysprositos (δυσπρόσιτος), meaning "hard to get". The element was not isolated in relatively pure form until after the development of ion exchange techniques by Frank Spedding at Iowa State University in the early 1950s. [9] [22]

Due to its role in permanent magnets used for wind turbines, it has been argued[ by whom? ] that dysprosium will be one of the main objects of geopolitical competition in a world running on renewable energy. But this perspective has been criticised for failing to recognise that most wind turbines do not use permanent magnets and for underestimating the power of economic incentives for expanded production. [23] [24]

In 2021, Dy was turned into a 2-dimensional supersolid quantum gas. [25]

Occurrence

Xenotime Xenotimio1.jpeg
Xenotime

While dysprosium is never encountered as a free element, it is found in many minerals, including xenotime, fergusonite, gadolinite, euxenite, polycrase, blomstrandine, monazite and bastnäsite, often with erbium and holmium or other rare earth elements. No dysprosium-dominant mineral (that is, with dysprosium prevailing over other rare earths in the composition) has yet been found. [26]

In the high-yttrium version of these, dysprosium happens to be the most abundant of the heavy lanthanides, comprising up to 7–8% of the concentrate (as compared to about 65% for yttrium). [27] [28] The concentration of Dy in the Earth's crust is about 5.2 mg/kg and in sea water 0.9 ng/L. [14]

Production

Dysprosium is obtained primarily from monazite sand, a mixture of various phosphates. The metal is obtained as a by-product in the commercial extraction of yttrium. In isolating dysprosium, most of the unwanted metals can be removed magnetically or by a flotation process. Dysprosium can then be separated from other rare earth metals by an ion exchange displacement process. The resulting dysprosium ions can then react with either fluorine or chlorine to form dysprosium fluoride, DyF3, or dysprosium chloride, DyCl3. These compounds can be reduced using either calcium or lithium metals in the following reactions: [15]

3 Ca + 2 DyF3 → 2 Dy + 3 CaF2
3 Li + DyCl3 → Dy + 3 LiCl

The components are placed in a tantalum crucible and fired in a helium atmosphere. As the reaction progresses, the resulting halide compounds and molten dysprosium separate due to differences in density. When the mixture cools, the dysprosium can be cut away from the impurities. [15]

About 100 tonnes of dysprosium are produced worldwide each year, [29] with 99% of that total produced in China. [30] Dysprosium prices have climbed nearly twentyfold, from $7 per pound in 2003, to $130 a pound in late 2010. [30] The price increased to $1,400/kg in 2011 but fell to $240 in 2015, largely due to illegal production in China which circumvented government restrictions. [31]

Currently, most dysprosium is being obtained from the ion-adsorption clay ores of southern China. [32] As of November 2018 the Browns Range Project pilot plant, 160 km south east of Halls Creek, Western Australia, is producing 50 tonnes (49 long tons) per annum. [33] [34]

According to the United States Department of Energy, the wide range of its current and projected uses, together with the lack of any immediately suitable replacement, makes dysprosium the single most critical element for emerging clean energy technologies; even their most conservative projections predicted a shortfall of dysprosium before 2015. [35] As of late 2015, there is a nascent rare earth (including dysprosium) extraction industry in Australia. [36]

Applications

Dysprosium is used, in conjunction with vanadium and other elements, in making laser materials and commercial lighting. Because of dysprosium's high thermal-neutron absorption cross-section, dysprosium-oxide–nickel cermets are used in neutron-absorbing control rods in nuclear reactors. [9] [37] Dysprosium–cadmium chalcogenides are sources of infrared radiation, which is useful for studying chemical reactions. [8] Because dysprosium and its compounds are highly susceptible to magnetization, they are employed in various data-storage applications, such as in hard disks. [38] Dysprosium is increasingly in demand for the permanent magnets used in electric-car motors and wind-turbine generators. [39]

Neodymium–iron–boron magnets can have up to 6% of the neodymium substituted by dysprosium [40] to raise the coercivity for demanding applications, such as drive motors for electric vehicles and generators for wind turbines. This substitution would require up to 100 grams of dysprosium per electric car produced. Based on Toyota's projected 2 million units per year, the use of dysprosium in applications such as this would quickly exhaust its available supply. [41] The dysprosium substitution may also be useful in other applications because it improves the corrosion resistance of the magnets. [42]

Dysprosium is one of the components of Terfenol-D, along with iron and terbium. Terfenol-D has the highest room-temperature magnetostriction of any known material, [43] which is employed in transducers, wide-band mechanical resonators, [44] and high-precision liquid-fuel injectors. [45]

Dysprosium is used in dosimeters for measuring ionizing radiation. Crystals of calcium sulfate or calcium fluoride are doped with dysprosium. When these crystals are exposed to radiation, the dysprosium atoms become excited and luminescent. The luminescence can be measured to determine the degree of exposure to which the dosimeter has been subjected. [9]

Nanofibers of dysprosium compounds have high strength and a large surface area. Therefore, they can be used to reinforce other materials and act as a catalyst. Fibers of dysprosium oxide fluoride can be produced by heating an aqueous solution of DyBr3 and NaF to 450 °C at 450  bars for 17 hours. This material is remarkably robust, surviving over 100 hours in various aqueous solutions at temperatures exceeding 400 °C without redissolving or aggregating. [46] [47] [48] Additionally, dysprosium has been used to create a two dimensional supersolid in a laboratory environment. Supersolids are expected to exhibit unusual properties, including superfluidity. [49]

Dysprosium iodide and dysprosium bromide are used in high-intensity metal-halide lamps. These compounds dissociate near the hot center of the lamp, releasing isolated dysprosium atoms. The latter re-emit light in the green and red part of the spectrum, thereby effectively producing bright light. [9] [50]

Several paramagnetic crystal salts of dysprosium (dysprosium gallium garnet, DGG; dysprosium aluminium garnet, DAG; dysprosium iron garnet, DyIG) are used in adiabatic demagnetization refrigerators. [51] [52]

The trivalent dysprosium ion (Dy3+) has been studied due to its downshifting luminescence properties. Dy-doped yttrium aluminium garnet (Dy:YAG) excited in the ultraviolet region of the electromagnetic spectrum results in the emission of photons of longer wavelength in the visible region. This idea is the basis for a new generation of UV-pumped white light-emitting diodes. [53]

The stable isotopes of dysprosium have been laser cooled and confined in magneto-optical traps [54] for quantum physics experiments. The first Bose and Fermi quantum degenerate gases of an open shell lanthanide were created with dysprosium. [55] [56] Because dysprosium is highly magnetic—indeed it is the most magnetic fermionic element and nearly tied with terbium for most magnetic bosonic atom [57] —such gases serve as the basis for quantum simulation with strongly dipolar atoms. [58]

Precautions

Like many powders, dysprosium powder may present an explosion hazard when mixed with air and when an ignition source is present. Thin foils of the substance can also be ignited by sparks or by static electricity. Dysprosium fires cannot be extinguished with water. It can react with water to produce flammable hydrogen gas. [59] Dysprosium chloride fires can be extinguished with water. [60] Dysprosium fluoride and dysprosium oxide are non-flammable. [61] [62] Dysprosium nitrate, Dy(NO3)3, is a strong oxidizing agent and readily ignites on contact with organic substances. [10]

Soluble dysprosium salts, such as dysprosium chloride and dysprosium nitrate are mildly toxic when ingested. Based on the toxicity of dysprosium chloride to mice, it is estimated that the ingestion of 500 grams or more could be fatal to a human (c.f. lethal dose of 300 grams of common table salt for a 100 kilogram human). The insoluble salts are non-toxic. [9]

Related Research Articles

The actinide or actinoid series encompasses at least the 14 metallic chemical elements in the 5f series, with atomic numbers from 89 to 102, actinium through nobelium. The actinide series derives its name from the first element in the series, actinium. The informal chemical symbol An is used in general discussions of actinide chemistry to refer to any actinide.

<span class="mw-page-title-main">Europium</span> Chemical element, symbol Eu and atomic number 63

Europium is a chemical element; it has symbol Eu and atomic number 63. Europium is a silvery-white metal of the lanthanide series that reacts readily with air to form a dark oxide coating. It is the most chemically reactive, least dense, and softest of the lanthanide elements. It is soft enough to be cut with a knife. Europium was isolated in 1901 and named after the continent of Europe. Europium usually assumes the oxidation state +3, like other members of the lanthanide series, but compounds having oxidation state +2 are also common. All europium compounds with oxidation state +2 are slightly reducing. Europium has no significant biological role and is relatively non-toxic compared to other heavy metals. Most applications of europium exploit the phosphorescence of europium compounds. Europium is one of the rarest of the rare-earth elements on Earth.

<span class="mw-page-title-main">Erbium</span> Chemical element, symbol Er and atomic number 68

Erbium is a chemical element; it has symbol Er and atomic number 68. A silvery-white solid metal when artificially isolated, natural erbium is always found in chemical combination with other elements. It is a lanthanide, a rare-earth element, originally found in the gadolinite mine in Ytterby, Sweden, which is the source of the element's name.

<span class="mw-page-title-main">Gadolinium</span> Chemical element, symbol Gd and atomic number 64

Gadolinium is a chemical element; it has symbol Gd and atomic number 64. Gadolinium is a silvery-white metal when oxidation is removed. It is a malleable and ductile rare-earth element. Gadolinium reacts with atmospheric oxygen or moisture slowly to form a black coating. Gadolinium below its Curie point of 20 °C (68 °F) is ferromagnetic, with an attraction to a magnetic field higher than that of nickel. Above this temperature it is the most paramagnetic element. It is found in nature only in an oxidized form. When separated, it usually has impurities of the other rare-earths because of their similar chemical properties.

<span class="mw-page-title-main">Holmium</span> Chemical element, symbol Ho and atomic number 67

Holmium is a chemical element; it has symbol Ho and atomic number 67. It is a rare-earth element and the eleventh member of the lanthanide series. It is a relatively soft, silvery, fairly corrosion-resistant and malleable metal. Like many other lanthanides, holmium is too reactive to be found in native form, as pure holmium slowly forms a yellowish oxide coating when exposed to air. When isolated, holmium is relatively stable in dry air at room temperature. However, it reacts with water and corrodes readily, and also burns in air when heated.

<span class="mw-page-title-main">Lanthanum</span> Chemical element, symbol La and atomic number 57

Lanthanum is a chemical element; it has symbol La and atomic number 57. It is a soft, ductile, silvery-white metal that tarnishes slowly when exposed to air. It is the eponym of the lanthanide series, a group of 15 similar elements between lanthanum and lutetium in the periodic table, of which lanthanum is the first and the prototype. Lanthanum is traditionally counted among the rare earth elements. Like most other rare earth elements, the usual oxidation state is +3, although some compounds are known with an oxidation state of +2. Lanthanum has no biological role in humans but is essential to some bacteria. It is not particularly toxic to humans but does show some antimicrobial activity.

<span class="mw-page-title-main">Lutetium</span> Chemical element, symbol Lu and atomic number 71

Lutetium is a chemical element; it has symbol Lu and atomic number 71. It is a silvery white metal, which resists corrosion in dry air, but not in moist air. Lutetium is the last element in the lanthanide series, and it is traditionally counted among the rare earth elements; it can also be classified as the first element of the 6th-period transition metals.

The lanthanide or lanthanoid series of chemical elements comprises at least the 14 metallic chemical elements with atomic numbers 57–70, from lanthanum through ytterbium. In the periodic table, they fill the 4f orbitals. Lutetium is also sometimes considered a lanthanide, despite being a d-block element and a transition metal.

<span class="mw-page-title-main">Neodymium</span> Chemical element, symbol Nd and atomic number 60

Neodymium is a chemical element; it has symbol Nd and atomic number 60. It is the fourth member of the lanthanide series and is considered to be one of the rare-earth metals. It is a hard, slightly malleable, silvery metal that quickly tarnishes in air and moisture. When oxidized, neodymium reacts quickly producing pink, purple/blue and yellow compounds in the +2, +3 and +4 oxidation states. It is generally regarded as having one of the most complex spectra of the elements. Neodymium was discovered in 1885 by the Austrian chemist Carl Auer von Welsbach, who also discovered praseodymium. It is present in significant quantities in the minerals monazite and bastnäsite. Neodymium is not found naturally in metallic form or unmixed with other lanthanides, and it is usually refined for general use. Neodymium is fairly common—about as common as cobalt, nickel, or copper—and is widely distributed in the Earth's crust. Most of the world's commercial neodymium is mined in China, as is the case with many other rare-earth metals.

<span class="mw-page-title-main">Promethium</span> Chemical element, symbol Pm and atomic number 61

Promethium is a chemical element; it has symbol Pm and atomic number 61. All of its isotopes are radioactive; it is extremely rare, with only about 500–600 grams naturally occurring in Earth's crust at any given time. Promethium is one of only two radioactive elements that are followed in the periodic table by elements with stable forms, the other being technetium. Chemically, promethium is a lanthanide. Promethium shows only one stable oxidation state of +3.

<span class="mw-page-title-main">Scandium</span> Chemical element, symbol Sc and atomic number 21

Scandium is a chemical element; it has symbol Sc and atomic number 21. It is a silvery-white metallic d-block element. Historically, it has been classified as a rare-earth element, together with yttrium and the lanthanides. It was discovered in 1879 by spectral analysis of the minerals euxenite and gadolinite from Scandinavia.

<span class="mw-page-title-main">Terbium</span> Chemical element, symbol Tb and atomic number 65

Terbium is a chemical element; it has the symbol Tb and atomic number 65. It is a silvery-white, rare earth metal that is malleable, and ductile. The ninth member of the lanthanide series, terbium is a fairly electropositive metal that reacts with water, evolving hydrogen gas. Terbium is never found in nature as a free element, but it is contained in many minerals, including cerite, gadolinite, monazite, xenotime and euxenite.

<span class="mw-page-title-main">Thulium</span> Chemical element, symbol Tm and atomic number 69

Thulium is a chemical element; it has symbol Tm and atomic number 69. It is the thirteenth and third-last element in the lanthanide series. Like the other lanthanides, the most common oxidation state is +3, seen in its oxide, halides and other compounds; however, the +2 oxidation state can also be stable. In aqueous solution, like compounds of other late lanthanides, soluble thulium compounds form coordination complexes with nine water molecules.

<span class="mw-page-title-main">Ytterbium</span> Chemical element, symbol Yb and atomic number 70

Ytterbium is a chemical element; it has symbol Yb and atomic number 70. It is a metal, the fourteenth and penultimate element in the lanthanide series, which is the basis of the relative stability of its +2 oxidation state. Like the other lanthanides, its most common oxidation state is +3, as in its oxide, halides, and other compounds. In aqueous solution, like compounds of other late lanthanides, soluble ytterbium compounds form complexes with nine water molecules. Because of its closed-shell electron configuration, its density, melting point and boiling point are much lower than those of most other lanthanides.

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

<span class="mw-page-title-main">Praseodymium</span> Chemical element, symbol Pr and atomic number 59

Praseodymium is a chemical element; it has symbol Pr and the atomic number 59. It is the third member of the lanthanide series and is considered one of the rare-earth metals. It is a soft, silvery, malleable and ductile metal, valued for its magnetic, electrical, chemical, and optical properties. It is too reactive to be found in native form, and pure praseodymium metal slowly develops a green oxide coating when exposed to air.

<span class="mw-page-title-main">Group 3 element</span> Group of chemical elements

Group 3 is the first group of transition metals in the periodic table. This group is closely related to the rare-earth elements. It contains the four elements scandium (Sc), yttrium (Y), lutetium (Lu), and lawrencium (Lr). The group is also called the scandium group or scandium family after its lightest member.

<span class="mw-page-title-main">Yttrium</span> Chemical element, symbol Y and atomic number 39

Yttrium is a chemical element; it has symbol Y and atomic number 39. It is a silvery-metallic transition metal chemically similar to the lanthanides and has often been classified as a "rare-earth element". Yttrium is almost always found in combination with lanthanide elements in rare-earth minerals and is never found in nature as a free element. 89Y is the only stable isotope and the only isotope found in the Earth's crust.

<span class="mw-page-title-main">Cerium</span> Chemical element, symbol Ce and atomic number 58

Cerium is a chemical element; it has symbol Ce and atomic number 58. Cerium is a soft, ductile, and silvery-white metal that tarnishes when exposed to air. Cerium is the second element in the lanthanide series, and while it often shows the oxidation state of +3 characteristic of the series, it also has a stable +4 state that does not oxidize water. It is also considered one of the rare-earth elements. Cerium has no known biological role in humans but is not particularly toxic, except with intense or continued exposure.

<span class="mw-page-title-main">Terbium compounds</span> Chemical compounds with at least one terbium atom

Terbium compounds are compounds formed by the lanthanide metal terbium (Tb). Terbium generally exhibits the +3 oxidation state in these compounds, such as in TbCl3, Tb(NO3)3 and Tb(CH3COO)3. Compounds with terbium in the +4 oxidation state are also known, such as TbO2 and BaTbF6. Terbium can also form compounds in the 0, +1 and +2 oxidation states.

References

  1. "Standard Atomic Weights: Dysprosium". CIAAW. 2001.
  2. Prohaska, Thomas; Irrgeher, Johanna; Benefield, Jacqueline; Böhlke, John K.; Chesson, Lesley A.; Coplen, Tyler B.; Ding, Tiping; Dunn, Philip J. H.; Gröning, Manfred; Holden, Norman E.; Meijer, Harro A. J. (2022-05-04). "Standard atomic weights of the elements 2021 (IUPAC Technical Report)". Pure and Applied Chemistry. doi:10.1515/pac-2019-0603. ISSN   1365-3075.
  3. 1 2 3 4 Arblaster, John W. (2018). Selected Values of the Crystallographic Properties of Elements. Materials Park, Ohio: ASM International. ISBN   978-1-62708-155-9.
  4. Yttrium and all lanthanides except Ce and Pm have been observed in the oxidation state 0 in bis(1,3,5-tri-t-butylbenzene) complexes, see Cloke, F. Geoffrey N. (1993). "Zero Oxidation State Compounds of Scandium, Yttrium, and the Lanthanides". Chem. Soc. Rev. 22: 17–24. doi:10.1039/CS9932200017. and Arnold, Polly L.; Petrukhina, Marina A.; Bochenkov, Vladimir E.; Shabatina, Tatyana I.; Zagorskii, Vyacheslav V.; Cloke (2003-12-15). "Arene complexation of Sm, Eu, Tm and Yb atoms: a variable temperature spectroscopic investigation". Journal of Organometallic Chemistry. 688 (1–2): 49–55. doi:10.1016/j.jorganchem.2003.08.028.
  5. Weast, Robert (1984). CRC, Handbook of Chemistry and Physics. Boca Raton, Florida: Chemical Rubber Company Publishing. pp. E110. ISBN   0-8493-0464-4.
  6. Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S.; Audi, G. (2021). "The NUBASE2020 evaluation of nuclear properties" (PDF). Chinese Physics C. 45 (3): 030001. doi:10.1088/1674-1137/abddae.
  7. Chiera, Nadine Mariel; Dressler, Rugard; Sprung, Peter; Talip, Zeynep; Schumann, Dorothea (2022-05-28). "High precision half-life measurement of the extinct radio-lanthanide Dysprosium-154". Scientific Reports. 12 (1). Springer Science and Business Media LLC. doi:10.1038/s41598-022-12684-6. ISSN   2045-2322.
  8. 1 2 Lide, David R., ed. (2007–2008). "Dysprosium". CRC Handbook of Chemistry and Physics. Vol. 4. New York: CRC Press. p. 11. ISBN   978-0-8493-0488-0.
  9. 1 2 3 4 5 6 Emsley, John (2001). Nature's Building Blocks. Oxford: Oxford University Press. pp. 129–132. ISBN   978-0-19-850341-5.
  10. 1 2 3 Krebs, Robert E. (1998). "Dysprosium". The History and Use of our Earth's Chemical Elements. Greenwood Press. pp.  234–235. ISBN   978-0-313-30123-0.
  11. Jackson, Mike (2000). "Wherefore Gadolinium? Magnetism of the Rare Earths" (PDF). IRM Quarterly. 10 (3): 6. Archived from the original (PDF) on 2017-07-12. Retrieved 2009-05-03.
  12. Junyang Jin, Yaru Ni, Wenjuan Huang, Chunhua Lu, Zhongzi Xu (March 2013). "Controlled synthesis and characterization of large-scale, uniform sheet-shaped dysprosium hydroxide nanosquares by hydrothermal method". Journal of Alloys and Compounds. 553: 333–337. doi:10.1016/j.jallcom.2012.11.068. ISSN   0925-8388 . Retrieved 2018-06-13.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  13. "Chemical reactions of Dysprosium". Webelements. Retrieved 2012-08-16.
  14. 1 2 Patnaik, Pradyot (2003). Handbook of Inorganic Chemical Compounds. McGraw-Hill. pp. 289–290. ISBN   978-0-07-049439-8 . Retrieved 2009-06-06.
  15. 1 2 3 Heiserman, David L. (1992). Exploring Chemical Elements and their Compounds . TAB Books. pp.  236–238. ISBN   978-0-8306-3018-9.
  16. Perry, D. L. (1995). Handbook of Inorganic Compounds. CRC Press. pp. 152–154. ISBN   978-0-8493-8671-8.
  17. Jantsch, G.; Ohl, A. (1911). "Zur Kenntnis der Verbindungen des Dysprosiums". Berichte der Deutschen Chemischen Gesellschaft. 44 (2): 1274–1280. doi:10.1002/cber.19110440215.
  18. Vallina, B., Rodriguez-Blanco, J.D., Brown, A.P., Blanco, J.A. and Benning, L.G. (2013). "Amorphous dysprosium carbonate: characterization, stability and crystallization pathways". Journal of Nanoparticle Research. 15 (2): 1438. Bibcode:2013JNR....15.1438V. CiteSeerX   10.1.1.705.3019 . doi:10.1007/s11051-013-1438-3. S2CID   95924050.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  19. 1 2 3 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.
  20. DeKosky, Robert K. (1973). "Spectroscopy and the Elements in the Late Nineteenth Century: The Work of Sir William Crookes". The British Journal for the History of Science. 6 (4): 400–423. doi:10.1017/S0007087400012553. JSTOR   4025503. S2CID   146534210.
  21. de Boisbaudran, Paul Émile Lecoq (1886). "L'holmine (ou terre X de M Soret) contient au moins deux radicaux métallique (Holminia contains at least two metal)". Comptes Rendus (in French). 143: 1003–1006.
  22. Weeks, Mary Elvira (1956). The discovery of the elements (6th ed.). Easton, PA: Journal of Chemical Education.
  23. Overland, Indra (2019-03-01). "The geopolitics of renewable energy: Debunking four emerging myths". Energy Research & Social Science. 49: 36–40. doi: 10.1016/j.erss.2018.10.018 . ISSN   2214-6296.
  24. Klinger, Julie Michelle (2017). Rare earth frontiers : from terrestrial subsoils to lunar landscapes. Ithaca, NY: Cornell University Press. ISBN   978-1501714603. JSTOR   10.7591/j.ctt1w0dd6d.
  25. Norcia, Matthew A.; Politi, Claudia; Klaus, Lauritz; Poli, Elena; Sohmen, Maximilian; Mark, Manfred J.; Bisset, Russell N.; Santos, Luis; Ferlaino, Francesca (August 2021). "Two-dimensional supersolidity in a dipolar quantum gas". Nature. 596 (7872): 357–361. arXiv: 2102.05555 . Bibcode:2021Natur.596..357N. doi:10.1038/s41586-021-03725-7. ISSN   1476-4687. PMID   34408330. S2CID   231861397.
  26. Hudson Institute of Mineralogy (1993–2018). "Mindat.org". www.mindat.org. Retrieved 14 January 2018.
  27. Naumov, A. V. (2008). "Review of the World Market of Rare-Earth Metals". Russian Journal of Non-Ferrous Metals. 49 (1): 14–22. doi:10.1007/s11981-008-1004-6. S2CID   135730387.
  28. Gupta, C. K.; Krishnamurthy N. (2005). Extractive Metallurgy of Rare Earths. CRC Press. ISBN   978-0-415-33340-5.
  29. "Dysprosium (Dy) - Chemical properties, Health and Environmental effects". Lenntech Water treatment & air purification Holding B.V. 2008. Retrieved 2009-06-02.
  30. 1 2 Bradsher, Keith (December 29, 2010). "In China, Illegal Rare Earth Mines Face Crackdown". The New York Times.
  31. Rare Earths archive. United States Geological Survey . January 2016
  32. Bradsher, Keith (December 25, 2009). "Earth-Friendly Elements, Mined Destructively". The New York Times.
  33. Major, Tom (30 November 2018). "Rare earth mineral discovery set to make Australia a major player in electric vehicle supply chain". ABC News. Australian Broadcasting Corporation. Retrieved 30 November 2018.
  34. Brann, Matt (November 27, 2011). "Halls Creek turning into a hub for rare earths".
  35. New Scientist, 18 June 2011, p. 40
  36. Jasper, Clint (2015-09-22) Staring down a multitude of challenges, these Australian rare earth miners are confident they can break into the market. abc.net.au
  37. Amit, Sinha; Sharma, Beant Prakash (2005). "Development of Dysprosium Titanate Based Ceramics". Journal of the American Ceramic Society. 88 (4): 1064–1066. doi:10.1111/j.1551-2916.2005.00211.x.
  38. Lagowski, J. J., ed. (2004). Chemistry Foundations and Applications. Vol. 2. Thomson Gale. pp.  267–268. ISBN   978-0-02-865724-0.
  39. Bourzac, Katherine (19 April 2011). "The Rare Earth Crisis". MIT Technology Review. Retrieved 18 June 2016.
  40. Shi, Fang, X.; Shi, Y.; Jiles, D. C. (1998). "Modeling of magnetic properties of heat treated Dy-doped NdFeB particles bonded in isotropic and anisotropic arrangements". IEEE Transactions on Magnetics (Submitted manuscript). 34 (4): 1291–1293. Bibcode:1998ITM....34.1291F. doi:10.1109/20.706525.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  41. Campbell, Peter (February 2008). "Supply and Demand, Part 2". Princeton Electro-Technology, Inc. Archived from the original on June 4, 2008. Retrieved 2008-11-09.
  42. Yu, L. Q.; Wen, Y.; Yan, M. (2004). "Effects of Dy and Nb on the magnetic properties and corrosion resistance of sintered NdFeB". Journal of Magnetism and Magnetic Materials. 283 (2–3): 353–356. Bibcode:2004JMMM..283..353Y. doi:10.1016/j.jmmm.2004.06.006.
  43. "What is Terfenol-D?". ETREMA Products, Inc. 2003. Archived from the original on 2015-05-10. Retrieved 2008-11-06.
  44. Kellogg, Rick; Flatau, Alison (May 2004). "Wide Band Tunable Mechanical Resonator Employing the ΔE Effect of Terfenol-D". Journal of Intelligent Material Systems & Structures. 15 (5): 355–368. doi:10.1177/1045389X04040649. S2CID   110609960.
  45. Leavitt, Wendy (February 2000). "Take Terfenol-D and call me". Fleet Owner. 95 (2): 97. Retrieved 2008-11-06.
  46. "Supercritical Water Oxidation/Synthesis". Pacific Northwest National Laboratory. Archived from the original on 2008-04-20. Retrieved 2009-06-06.
  47. "Rare Earth Oxide Fluoride: Ceramic Nano-particles via a Hydrothermal Method". Pacific Northwest National Laboratory. Archived from the original on 2010-05-27. Retrieved 2009-06-06.{{cite web}}: CS1 maint: bot: original URL status unknown (link)
  48. Hoffman, M. M.; Young, J. S.; Fulton, J. L. (2000). "Unusual dysprosium ceramic nano-fiber growth in a supercritical aqueous solution". J. Mater. Sci. 35 (16): 4177. Bibcode:2000JMatS..35.4177H. doi:10.1023/A:1004875413406. S2CID   55710942.
  49. "Physicists give weird new phase of matter an extra dimension". Live Science. 18 August 2021. Retrieved 18 August 2021.
  50. Gray, Theodore (2009). The Elements. Black Dog and Leventhal Publishers. pp.  152–153. ISBN   978-1-57912-814-2.
  51. Milward, Steve et al. (2004). "Design, Manufacture and Test of an Adiabatic Demagnetization Refrigerator Magnet for use in Space". Archived 2013-10-04 at the Wayback Machine . University College London.
  52. Hepburn, Ian. "Adiabatic Demagnetization Refrigerator: A Practical Point of View". Archived 2013-10-04 at the Wayback Machine . Cryogenic Physics Group, Mullard Space Science Laboratory, University College London.
  53. Carreira, J. F. C. (2017). "YAG:Dy – Based single white light emitting phosphor produced by solution combustion synthesis". Journal of Luminescence. 183: 251–258. Bibcode:2017JLum..183..251C. doi:10.1016/j.jlumin.2016.11.017.
  54. Lu, M.; Youn, S.-H.; Lev, B. (2010). "Trapping Ultracold Dysprosium: A Highly Magnetic Gas for Dipolar Physics". Physical Review Letters. 104 (6): 063001. arXiv: 0912.0050 . Bibcode:2010PhRvL.104f3001L. doi:10.1103/physrevlett.104.063001. PMID   20366817. S2CID   7614035.
  55. Lu, M.; Burdick, N.; Youn, S.-H.; Lev, B. (2011). "Strongly Dipolar Bose–Einstein Condensate of Dysprosium". Physical Review Letters. 107 (19): 190401. arXiv: 1108.5993 . Bibcode:2011PhRvL.107s0401L. doi:10.1103/physrevlett.107.190401. PMID   22181585. S2CID   21945255.
  56. Lu, M.; Burdick, N.; Lev, B. (2012). "Quantum Degenerate Dipolar Fermi Gas". Physical Review Letters. 108 (21): 215301. arXiv: 1202.4444 . Bibcode:2012PhRvL.108u5301L. doi:10.1103/physrevlett.108.215301. PMID   23003275. S2CID   15650840.
  57. Martin, W C; Zalubas, R; Hagan, L (January 1978). "Atomic energy levels - the rare earth elements". OSTI.GOV. OSTI   6507735 . Retrieved 2023-03-11.
  58. Chomaz, L.; Ferrier-Barbut, I.; Ferlaino, F.; Laburthe-Tolra, B.; Lev, B.; Pfau, T. (2022). "Dipolar physics: a review of experiments with magnetic quantum gases". Rep. Prog. Phys. 86 (2): 026401. arXiv: 2201.02672 . doi:10.1088/1361-6633/aca814. PMID   36583342. S2CID   245837061.
  59. Dierks, Steve (January 2003). "Dysprosium". Material Safety Data Sheets. Electronic Space Products International. Archived from the original on 2015-09-22. Retrieved 2008-10-20.
  60. Dierks, Steve (January 1995). "Dysprosium Chloride". Material Safety Data Sheets. Electronic Space Products International. Archived from the original on 2015-09-22. Retrieved 2008-11-07.{{cite web}}: CS1 maint: bot: original URL status unknown (link)
  61. Dierks, Steve (December 1995). "Dysprosium Fluoride". Material Safety Data Sheets. Electronic Space Products International. Archived from the original on 2015-09-22. Retrieved 2008-11-07.{{cite web}}: CS1 maint: bot: original URL status unknown (link)
  62. Dierks, Steve (November 1988). "Dysprosium Oxide". Material Safety Data Sheets. Electronic Space Products International. Archived from the original on 2015-09-22. Retrieved 2008-11-07.{{cite web}}: CS1 maint: bot: original URL status unknown (link)