Ytterbium

Last updated
Ytterbium, 70Yb
Ytterbium-3.jpg
Ytterbium
Pronunciation /ɪˈtɜːrbiəm/ (ih-TUR-bee-əm)
Appearancesilvery white; with a pale yellow tint [1]
Standard atomic weight Ar°(Yb)
Ytterbium 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


Yb

No
thuliumytterbiumlutetium
Atomic number (Z)70
Group f-block groups (no number)
Period period 6
Block   f-block
Electron configuration [ Xe ] 4f14 6s2
Electrons per shell2, 8, 18, 32, 8, 2
Physical properties
Phase at  STP solid
Melting point 1097  K (824 °C,1515 °F)
Boiling point 1469 K(1196 °C,2185 °F)
Density (at 20° C)6.967 g/cm3 [4]
when liquid (at  m.p.)6.21 g/cm3
Heat of fusion 7.66  kJ/mol
Heat of vaporization 129 kJ/mol
Molar heat capacity 26.74 J/(mol·K)
Vapor pressure
P (Pa)1101001 k10 k100 k
at T (K)7368139101047(1266)(1465)
Atomic properties
Oxidation states 0, [5] +1, [6] +2, +3 (a  basic oxide)
Electronegativity Pauling scale: 1.1(?)
Ionization energies
  • 1st: 603.4 kJ/mol
  • 2nd: 1174.8 kJ/mol
  • 3rd: 2417 kJ/mol
Atomic radius empirical:176  pm
Covalent radius 187±8 pm
Ytterbium spectrum visible.png
Spectral lines of ytterbium
Other properties
Natural occurrence primordial
Crystal structure face-centered cubic (fcc)(cF4)
Lattice constants
Cubic-face-centered.svg
a = 548.46 pm (at 20 °C) [4]
Thermal expansion 24.31×10−6/K (at 20 °C) [4]
Thermal conductivity 38.5 W/(m⋅K)
Electrical resistivity β, poly: 0.250 µΩ⋅m(at r.t.)
Magnetic ordering paramagnetic
Molar magnetic susceptibility +249.0×10−6 cm3/mol(2928 K) [7]
Young's modulus β form: 23.9 GPa
Shear modulus β form: 9.9 GPa
Bulk modulus β form: 30.5 GPa
Speed of sound thin rod1590 m/s(at 20 °C)
Poisson ratio β form: 0.207
Vickers hardness 205–250 MPa
Brinell hardness 340–440 MPa
CAS Number 7440-64-4
History
Namingafter Ytterby (Sweden), where it was mined
Discovery Jean Charles Galissard de Marignac (1878)
First isolation Carl Auer von Welsbach (1906)
Isotopes of ytterbium
Main isotopes [8] Decay
abun­dance half-life (t1/2) mode pro­duct
166Yb synth 56.7 h ε 166Tm
168Yb0.126% stable
169Ybsynth32.026 dε 169Tm
170Yb3.02%stable
171Yb14.2%stable
172Yb21.8%stable
173Yb16.1%stable
174Yb31.9%stable
175Ybsynth4.185 d β 175Lu
176Yb12.9%stable
177Ybsynth1.911 hβ 177Lu
Symbol category class.svg  Category: Ytterbium
| references

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.

Contents

In 1878, Swiss chemist Jean Charles Galissard de Marignac separated from the rare earth "erbia" (another independent component) which he called "ytterbia", for Ytterby, the village in Sweden near where he found the new component of erbium. He suspected that ytterbia was a compound of a new element that he called "ytterbium" (in total, four elements were named after the village, the others being yttrium, terbium, and erbium). In 1907, the new earth "lutecia" was separated from ytterbia, from which the element "lutecium" (now lutetium) was extracted by Georges Urbain, Carl Auer von Welsbach, and Charles James. After some discussion, Marignac's name "ytterbium" was retained. A relatively pure sample of the metal was not obtained until 1953. At present, ytterbium is mainly used as a dopant of stainless steel or active laser media, and less often as a gamma ray source.

Natural ytterbium is a mixture of seven stable isotopes, which altogether are present at concentrations of 0.3 parts per million. This element is mined in China, the United States, Brazil, and India in form of the minerals monazite, euxenite, and xenotime. The ytterbium concentration is low because it is found only among many other rare-earth elements; moreover, it is among the least abundant. Once extracted and prepared, ytterbium is somewhat hazardous as an eye and skin irritant. The metal is a fire and explosion hazard.

Characteristics

Physical properties

Ytterbium is a soft, malleable and ductile chemical element. In color when freshly prepared, it is less golden than cesium, but, more golden in color than just a "yellow-cast" as in metals like iridium. It is a rare-earth element, and it is readily dissolved by the strong mineral acids. [9]

Ytterbium has three allotropes labeled by the Greek letters alpha, beta and gamma. Their transformation temperatures are −13 °C and 795 °C, [9] although the exact transformation temperature depends on the pressure and stress. [10] 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. [9] The alpha allotrope (6.903 g/cm3) has a hexagonal crystalline structure and is stable at low temperatures. [11] 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). [9] [12]

In contrast with the other rare-earth metals, which usually have antiferromagnetic and/or ferromagnetic properties at low temperatures, ytterbium is paramagnetic at temperatures above 1.0 kelvin. [13] However, the alpha allotrope is diamagnetic. [10] With a melting point of 824 °C and a boiling point of 1196 °C, ytterbium has the smallest liquid range of all the metals. [9]

Contrary to most other lanthanides, which have a close-packed hexagonal lattice, ytterbium crystallizes in the face-centered cubic system. Ytterbium has a density of 6.973 g/cm3, which is significantly lower than those of the neighboring lanthanides, thulium (9.32 g/cm3) and lutetium (9.841 g/cm3). Its melting and boiling points are also significantly lower than those of thulium and lutetium. This is due to the closed-shell electron configuration of ytterbium ([Xe] 4f14 6s2), which causes only the two 6s electrons to be available for metallic bonding (in contrast to the other lanthanides where three electrons are available) and increases ytterbium's metallic radius. [11]

Chemical properties

Ytterbium metal tarnishes slowly in air, taking on a golden or brown hue. Finely dispersed ytterbium readily oxidizes in air and under oxygen. Mixtures of powdered ytterbium with polytetrafluoroethylene or hexachloroethane burn with a luminous emerald-green flame. [14] Ytterbium reacts with hydrogen to form various non-stoichiometric hydrides. Ytterbium dissolves slowly in water, but quickly in acids, liberating hydrogen gas. [11]

Ytterbium is quite electropositive, and it reacts slowly with cold water and quite quickly with hot water to form ytterbium(III) hydroxide: [15]

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

Ytterbium reacts with all the halogens: [15]

2 Yb (s) + 3 F2 (g) → 2 YbF3 (s) [white]
2 Yb (s) + 3 Cl2 (g) → 2 YbCl3 (s) [white]
2 Yb (s) + 3 Br2 (l) → 2 YbBr3 (s) [white]
2 Yb (s) + 3 I2 (s) → 2 YbI3 (s) [white]

The ytterbium(III) ion absorbs light in the near infrared range of wavelengths, but not in visible light, so ytterbia, Yb2O3, is white in color and the salts of ytterbium are also colorless. Ytterbium dissolves readily in dilute sulfuric acid to form solutions that contain the colorless Yb(III) ions, which exist as nonahydrate complexes: [15]

2 Yb (s) + 3 H2SO4 (aq) + 18 H
2
O
(l) → 2 [Yb(H2O)9]3+ (aq) + 3 SO2−
4
(aq) + 3 H2 (g)

Yb(II) vs. Yb(III)

Although usually trivalent, ytterbium readily forms divalent compounds. This behavior is unusual for lanthanides, which almost exclusively form compounds with an oxidation state of +3. The +2 state has a valence electron configuration of 4f14 because the fully filled f-shell gives more stability. The yellow-green ytterbium(II) ion is a very strong reducing agent and decomposes water, releasing hydrogen gas, and thus only the colorless ytterbium(III) ion occurs in aqueous solution. Samarium and thulium also behave this way in the +2 state, but europium(II) is stable in aqueous solution. Ytterbium metal behaves similarly to europium metal and the alkaline earth metals, dissolving in ammonia to form blue electride salts. [11]

Isotopes

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). Thirty-two 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). [16] [8]

The isotopes of ytterbium range from 149Yb to 187Yb. [8] [17] 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. [16] [8]

Occurrence

Euxenite Euxenite - Vegusdal, Norvegia 01.jpg
Euxenite

Ytterbium is found with other rare-earth elements in several rare minerals. It is most often recovered commercially from monazite sand (0.03% ytterbium). The element is also found in euxenite and xenotime. The main mining areas are China, the United States, Brazil, India, Sri Lanka, and Australia. Reserves of ytterbium are estimated as one million tonnes. Ytterbium is normally difficult to separate from other rare earths, but ion-exchange and solvent extraction techniques developed in the mid- to late 20th century have simplified separation. Compounds of ytterbium are rare and have not yet been well characterized. The abundance of ytterbium in the Earth's crust is about 3 mg/kg. [12]

As an even-numbered lanthanide, in accordance with the Oddo-Harkins rule, ytterbium is significantly more abundant than its immediate neighbors, thulium and lutetium, which occur in the same concentrate at levels of about 0.5% each. The world production of ytterbium is only about 50 tonnes per year, reflecting that it has few commercial applications. [12] Microscopic traces of ytterbium are used as a dopant in the Yb:YAG laser, a solid-state laser in which ytterbium is the element that undergoes stimulated emission of electromagnetic radiation. [18]

Ytterbium is often the most common substitute in yttrium minerals. In very few known cases/occurrences ytterbium prevails over yttrium, as, e.g., in xenotime-(Yb). A report of native ytterbium from the Moon's regolith is known. [19]

Production

It is relatively difficult to separate ytterbium from other lanthanides due to its similar properties. As a result, the process is somewhat long. First, minerals such as monazite or xenotime are dissolved into various acids, such as sulfuric acid. Ytterbium can then be separated from other lanthanides by ion exchange, as can other lanthanides. The solution is then applied to a resin, which different lanthanides bind in different matters. This is then dissolved using complexing agents, and due to the different types of bonding exhibited by the different lanthanides, it is possible to isolate the compounds. [20] [21]

Ytterbium is separated from other rare earths either by ion exchange or by reduction with sodium amalgam. In the latter method, a buffered acidic solution of trivalent rare earths is treated with molten sodium-mercury alloy, which reduces and dissolves Yb3+. The alloy is treated with hydrochloric acid. The metal is extracted from the solution as oxalate and converted to oxide by heating. The oxide is reduced to metal by heating with lanthanum, aluminium, cerium or zirconium in high vacuum. The metal is purified by sublimation and collected over a condensed plate. [22]

Compounds

Ytterbium(III) oxide Ytterbium(III) oxide.jpg
Ytterbium(III) oxide

The chemical behavior of ytterbium is similar to that of the rest of the lanthanides. Most ytterbium compounds are found in the +3 oxidation state, and its salts in this oxidation state are nearly colorless. Like europium, samarium, and thulium, the trihalides of ytterbium can be reduced to the dihalides by hydrogen, zinc dust, or by the addition of metallic ytterbium. [11] 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). [11]

Halides

Crystal structure of ytterbium(III) oxide Kristallstruktur Lanthanoid-C-Typ.png
Crystal structure of ytterbium(III) oxide

Ytterbium forms both dihalides and trihalides with the halogens fluorine, chlorine, bromine, and iodine. The dihalides are susceptible to oxidation to the trihalides at room temperature and disproportionate to the trihalides and metallic ytterbium at high temperature: [11]

3 YbX2 → 2 YbX3 + Yb (X = F, Cl, Br, I)

Some ytterbium halides are used as reagents in organic synthesis. For example, ytterbium(III) chloride (YbCl3) is a Lewis acid and can be used as a catalyst in the Aldol [23] and Diels–Alder reactions. [24] Ytterbium(II) iodide (YbI2) may be used, like samarium(II) iodide, as a reducing agent for coupling reactions. [25] Ytterbium(III) fluoride (YbF3) is used as an inert and non-toxic tooth filling as it continuously releases fluoride ions, which are good for dental health, and is also a good X-ray contrast agent. [26]

Oxides

Ytterbium reacts with oxygen to form ytterbium(III) oxide (Yb2O3), which crystallizes in the "rare-earth C-type sesquioxide" structure which is related to the fluorite structure with one quarter of the anions removed, leading to ytterbium atoms in two different six coordinate (non-octahedral) environments. [27] Ytterbium(III) oxide can be reduced to ytterbium(II) oxide (YbO) with elemental ytterbium, which crystallizes in the same structure as sodium chloride. [11]

Borides

Ytterbium dodecaboride (YbB12) is a crystalline material that has been studied to understand various electronic and structural properties of many chemically related substances. It is a Kondo insulator. [28] It is a quantum material; under normal conditions, the interior of the bulk crystal is an insulator whereas the surface is highly conductive. [29] Among the rare earth elements, ytterbium is one of the few that can form a stable dodecaboride, a property attributed to its comparatively small atomic radius. [30]

History

Jean Charles Galissard de Marignac Galissard de Marignac.jpg
Jean Charles Galissard de Marignac

Ytterbium was discovered by the Swiss chemist Jean Charles Galissard de Marignac in the year 1878. While examining samples of gadolinite, Marignac found a new component in the earth then known as erbia, and he named it ytterbia, for Ytterby, the Swedish village near where he found the new component of erbium. Marignac suspected that ytterbia was a compound of a new element that he called "ytterbium". [12] [26] [31] [32] [33]

In 1907, the French chemist Georges Urbain separated Marignac's ytterbia into two components: neoytterbia and lutecia. Neoytterbia later became known as the element ytterbium, and lutecia became known as the element lutetium. The Austrian chemist Carl Auer von Welsbach independently isolated these elements from ytterbia at about the same time, but he called them aldebaranium and cassiopeium; [12] the American chemist Charles James also independently isolated these elements at about the same time. [34] Urbain and Welsbach accused each other of publishing results based on the other party. [35] [36] [37] The Commission on Atomic Mass, consisting of Frank Wigglesworth Clarke, Wilhelm Ostwald, and Georges Urbain, which was then responsible for the attribution of new element names, settled the dispute in 1909 by granting priority to Urbain and adopting his names as official ones, based on the fact that the separation of lutetium from Marignac's ytterbium was first described by Urbain. [35] After Urbain's names were recognized, neoytterbium was reverted to ytterbium.

The chemical and physical properties of ytterbium could not be determined with any precision until 1953, when the first nearly pure ytterbium metal was produced by using ion-exchange processes. [12] The price of ytterbium was relatively stable between 1953 and 1998 at about US$1,000/kg. [38]

Applications

Source of gamma rays

The 169Yb isotope (with a half-life of 32 days), which is created along with the short-lived 175Yb isotope (half-life 4.2 days) by neutron activation during the irradiation of ytterbium in nuclear reactors, has been used as a radiation source in portable X-ray machines. Like X-rays, the gamma rays emitted by the source pass through soft tissues of the body, but are blocked by bones and other dense materials. Thus, small 169Yb samples (which emit gamma rays) act like tiny X-ray machines useful for radiography of small objects. Experiments show that radiographs taken with a 169Yb source are roughly equivalent to those taken with X-rays having energies between 250 and 350 keV. 169Yb is also used in nuclear medicine. [39]

High-stability atomic clocks

In 2013, ytterbium clocks held the record for stability with ticks stable to within less than two parts in 1 quintillion (2×10−18). [40] These clocks developed at the National Institute of Standards and Technology (NIST) rely on about 10,000 ytterbium atoms laser-cooled to 10 microkelvin (10 millionths of a degree above absolute zero) and trapped in an optical lattice—a series of pancake-shaped wells made of laser light. Another laser that "ticks" 518 trillion times per second (518 THz) provokes a transition between two energy levels in the atoms. The large number of atoms is key to the clocks' high stability.

Visible light waves oscillate faster than microwaves, hence optical clocks can be more precise than caesium atomic clocks. The Physikalisch-Technische Bundesanstalt is working on several such optical clocks. The model with one single ytterbium ion caught in an ion trap is highly accurate. The optical clock based on it is exact to 17 digits after the decimal point. [41]

A pair of experimental atomic clocks based on ytterbium atoms at the National Institute of Standards and Technology has set a record for stability. NIST physicists reported in the August 22, 2013 issue of Science Express that the ytterbium clocks' ticks are stable to within less than two parts in 1 quintillion (1 followed by 18 zeros), roughly 10 times better than the previous best published results for other atomic clocks. The clocks would be accurate within a second for a period comparable to the age of the universe. [42]

Doping of stainless steel

Ytterbium can also be used as a dopant to help improve the grain refinement, strength, and other mechanical properties of stainless steel. Some ytterbium alloys have rarely been used in dentistry. [9] [12]

Ytterbium as dopant of active media

The Yb3+ ion is used as a doping material in active laser media, specifically in solid state lasers and double clad fiber lasers. Ytterbium lasers are highly efficient, have long lifetimes and can generate short pulses; ytterbium can also easily be incorporated into the material used to make the laser. [43] Ytterbium lasers commonly radiate in the 1.03–1.12  µm band being optically pumped at wavelength 900 nm–1 µm, dependently on the host and application. The small quantum defect makes ytterbium a prospective dopant for efficient lasers and power scaling. [44]

The kinetic of excitations in ytterbium-doped materials is simple and can be described within the concept of effective cross-sections; for most ytterbium-doped laser materials (as for many other optically pumped gain media), the McCumber relation holds, [45] [46] [47] although the application to the ytterbium-doped composite materials was under discussion. [48] [49]

Usually, low concentrations of ytterbium are used. At high concentrations, the ytterbium-doped materials show photodarkening [50] (glass fibers) or even a switch to broadband emission [51] (crystals and ceramics) instead of efficient laser action. This effect may be related with not only overheating, but also with conditions of charge compensation at high concentrations of ytterbium ions. [52]

Much progress has been made in the power scaling lasers and amplifiers produced with ytterbium (Yb) doped optical fibers. Power levels have increased from the 1 kW regimes due to the advancements in components as well as the Yb-doped fibers. Fabrication of Low NA, Large Mode Area fibers enable achievement of near perfect beam qualities (M2<1.1) at power levels of 1.5 kW to greater than 2 kW at ~1064 nm in a broadband configuration. [53] Ytterbium-doped LMA fibers also have the advantages of a larger mode field diameter, which negates the impacts of nonlinear effects such as stimulated Brillouin scattering and stimulated Raman scattering, which limit the achievement of higher power levels, and provide a distinct advantage over single mode ytterbium-doped fibers.

To achieve even higher power levels in ytterbium-based fiber systems, all factors of the fiber must be considered. These can be achieved only through optimization of all ytterbium fiber parameters, ranging from the core background losses to the geometrical properties, to reduce the splice losses within the cavity. Power scaling also requires optimization of matching passive fibers within the optical cavity. [54] The optimization of the ytterbium-doped glass itself through host glass modification of various dopants also plays a large part in reducing the background loss of the glass, improvements in slope efficiency of the fiber, and improved photodarkening performance, all of which contribute to increased power levels in 1 µm systems.

Ion qubits for quantum computing

The charged ion 171Yb+ is used by multiple academic groups and companies as the trapped-ion qubit for quantum computing. [55] [56] [57] Entangling gates, such as the Mølmer–Sørensen gate, have been achieved by addressing the ions with mode-locked pulse lasers. [58]

Others

Ytterbium metal increases its electrical resistivity when subjected to high stresses. This property is used in stress gauges to monitor ground deformations from earthquakes and explosions. [59]

Currently, ytterbium is being investigated as a possible replacement for magnesium in high density pyrotechnic payloads for kinematic infrared decoy flares. As ytterbium(III) oxide has a significantly higher emissivity in the infrared range than magnesium oxide, a higher radiant intensity is obtained with ytterbium-based payloads in comparison to those commonly based on magnesium/Teflon/Viton (MTV). [60]

Precautions

Although ytterbium is fairly stable chemically, it is stored in airtight containers and in an inert atmosphere such as a nitrogen-filled dry box to protect it from air and moisture. [61] All compounds of ytterbium are treated as highly toxic, although studies appear to indicate that the danger is minimal. However, ytterbium compounds cause irritation to human skin and eyes, and some might be teratogenic. [62] Metallic ytterbium dust can spontaneously combust, [63] and the resulting fumes are hazardous. Ytterbium fires cannot be extinguished using water, and only dry chemical class D fire extinguishers can extinguish the fires. [64]

Related Research Articles

<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">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 element in the lanthanide series of metals. It is the second-least abundant lanthanide in the Earth's crust, after radioactively unstable promethium. It is an easily workable metal with a bright silvery-gray luster. It is fairly soft and slowly tarnishes in air. Despite its high price and rarity, thulium is used as a dopant in solid-state lasers, and as the radiation source in some portable X-ray devices. It has no significant biological role and is not particularly toxic.

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">Ytterbium(III) oxide</span> Chemical compound

Ytterbium(III) oxide is the chemical compound with the formula Yb2O3. It is one of the more commonly encountered compounds of ytterbium. It occurs naturally in trace amounts in the mineral gadolinite. It was first isolated from this in 1878 by Jean Charles Galissard de Marignac.

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

Lutetium(III) oxide, a white solid, is a cubic compound of lutetium sometimes used in the preparation of specialty glasses. It is also called lutecia. It is a lanthanide oxide, also known as a rare earth.

<span class="mw-page-title-main">Georges Urbain</span>

Georges Urbain was a French chemist, a professor of the Sorbonne, a member of the Institut de France, and director of the Institute of Chemistry in Paris. Much of his work focused on the rare earths, isolating and separating elements such as europium and gadolinium, and studying their spectra, their magnetic properties and their atomic masses. He discovered the element lutetium. He also studied the efflorescence of saline hydrates.

<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.

A dopant is a small amount of a substance added to a material to alter its physical properties, such as electrical or optical properties. The amount of dopant is typically very low compared to the material being doped.

Ytterbium compounds are chemical compounds that contain the element ytterbium (Yb). The chemical behavior of ytterbium is similar to that of the rest of the lanthanides. Most ytterbium compounds are found in the +3 oxidation state, and its salts in this oxidation state are nearly colorless. Like europium, samarium, and thulium, the trihalides of ytterbium can be reduced to the dihalides by hydrogen, zinc dust, or by the addition of metallic ytterbium. 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).

References

  1. Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. p. 112. ISBN   978-0-08-037941-8.
  2. "Standard Atomic Weights: Ytterbium". CIAAW. 2015.
  3. 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.
  4. 1 2 3 Arblaster, John W. (2018). Selected Values of the Crystallographic Properties of Elements. Materials Park, Ohio: ASM International. ISBN   978-1-62708-155-9.
  5. 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.
  6. La(I), Pr(I), Tb(I), Tm(I), and Yb(I) have been observed in MB8 clusters; see Li, Wan-Lu; Chen, Teng-Teng; Chen, Wei-Jia; Li, Jun; Wang, Lai-Sheng (2021). "Monovalent lanthanide(I) in borozene complexes". Nature Communications. 12 (1): 6467. doi:10.1038/s41467-021-26785-9. PMC   8578558 . PMID   34753931.
  7. Weast, Robert (1984). CRC, Handbook of Chemistry and Physics. Boca Raton, Florida: Chemical Rubber Company Publishing. pp. E110. ISBN   0-8493-0464-4.
  8. 1 2 3 4 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.
  9. 1 2 3 4 5 6 Hammond, C. R. (2000). The Elements, in Handbook of Chemistry and Physics (81st ed.). CRC press. ISBN   978-0-8493-0481-1.
  10. 1 2 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. Bibcode:1970PhRvB...2.3911B. doi:10.1103/PhysRevB.2.3911.
  11. 1 2 3 4 5 6 7 8 Holleman, Arnold F.; Wiberg, Egon; Wiberg, Nils (1985). "Die Lanthanoide". Lehrbuch der Anorganischen Chemie (in German) (91–100 ed.). Walter de Gruyter. pp. 1265–1279. ISBN   978-3-11-007511-3.
  12. 1 2 3 4 5 6 7 Emsley, John (2003). Nature's building blocks: an A-Z guide to the elements . Oxford University Press. pp.  492–494. ISBN   978-0-19-850340-8.
  13. Jackson, M. (2000). "Magnetism of Rare Earth". The IRM quarterly 10(3): 1
  14. Koch, E. C.; Weiser, V.; Roth, E.; Knapp, S.; Kelzenberg, S. (2012). "Combustion of Ytterbium Metal". Propellants, Explosives, Pyrotechnics. 37: 9–11. doi:10.1002/prep.201100141.
  15. 1 2 3 "Chemical reactions of Ytterbium". Webelements. Retrieved 2009-06-06.
  16. 1 2 "Nucleonica: Universal Nuclide Chart". Nucleonica. 2007–2011. Retrieved July 22, 2011.
  17. Tarasov, O. B.; Gade, A.; Fukushima, K.; et al. (2024). "Observation of New Isotopes in the Fragmentation of 198Pt at FRIB". Physical Review Letters. 132 (072501). doi:10.1103/PhysRevLett.132.072501.
  18. Lacovara, P.; Choi, H. K.; Wang, C. A.; Aggarwal, R. L.; Fan, T. Y. (1991). "Room-Temperature Diode-Pumped Yb:YAG laser". Optics Letters. 16 (14): 1089–1091. Bibcode:1991OptL...16.1089L. doi:10.1364/OL.16.001089. PMID   19776885.
  19. Hudson Institute of Mineralogy (1993–2018). "Mindat.org". www.mindat.org. Retrieved 7 April 2018.
  20. Gelis, V. M.; Chuveleva, E. A.; Firsova, L. A.; Kozlitin, E. A.; Barabanov, I. R. (2005). "Optimization of Separation of Ytterbium and Lutetium by Displacement Complexing Chromatography". Russian Journal of Applied Chemistry. 78 (9): 1420. doi:10.1007/s11167-005-0530-6. S2CID   94642269.
  21. Hubicka, H.; Drobek, D. (1997). "Anion-Exchange Method for Separation of Ytterbium from Holmium and Erbium". Hydrometallurgy. 47 (1): 127–136. Bibcode:1997HydMe..47..127H. doi:10.1016/S0304-386X(97)00040-6.
  22. Patnaik, Pradyot (2003). Handbook of Inorganic Chemical Compounds. McGraw-Hill. pp. 973–975. ISBN   978-0-07-049439-8 . Retrieved 2009-06-06.
  23. Lou, S.; Westbrook, J. A.; Schaus, S. E. (2004). "Decarboxylative Aldol Reactions of Allyl β-Keto Esters via Heterobimetallic Catalysis". Journal of the American Chemical Society. 126 (37): 11440–11441. doi:10.1021/ja045981k. PMID   15366881.
  24. Fang, X.; Watkin, J. G.; Warner, B. P. (2000). "Ytterbium Trichloride-Catalyzed Allylation of Aldehydes with Allyltrimethylsilane". Tetrahedron Letters. 41 (4): 447. doi:10.1016/S0040-4039(99)02090-0.
  25. Girard, P.; Namy, J. L.; Kagan, H. B. (1980). "Divalent Lanthanide Derivatives in Organic Synthesis. 1. Mild Preparation of Samarium Iodide and Ytterbium Iodide and Their Use as Reducing or Coupling Agents". Journal of the American Chemical Society. 102 (8): 2693. doi:10.1021/ja00528a029.
  26. 1 2 Enghag, Per (2004). Encyclopedia of the elements: technical data, history, processing, applications. John Wiley & Sons, ISBN   978-3-527-30666-4, p. 448.
  27. Wells A.F. (1984) Structural Inorganic Chemistry 5th edition, Oxford Science Publications, ISBN   0-19-855370-6
  28. Al'tshuler, T. S.; Bresler, M. S. (2002). "On the nature of the energy gap in ytterbium dodecaboride YbB12". Physics of the Solid State. 44 (8): 1532–1535. Bibcode:2002PhSS...44.1532A. doi:10.1134/1.1501353. S2CID   120575196.
  29. Xiang, Z.; Kasahara, Y.; Asaba, T.; Lawson, B.; Tinsman, C.; Chen, Lu; Sugimoto, K.; Kawaguchi, S.; Sato, Y.; Li, G.; Yao, S.; Chen, Y. L.; Iga, F.; Singleton, John; Matsuda, Y.; Li, Lu (2018). "Quantum oscillations of electrical resistivity in an insulator". Science. 362 (6410): 65–69. arXiv: 1905.05140 . Bibcode:2018Sci...362...65X. doi:10.1126/science.aap9607. PMID   30166438. S2CID   206664739.
  30. La Placa, 1 S. J.; Noonan, D. (1963). "Ytterbium and terbium dodecaborides". Acta Crystallographica. 16 (11): 1182. doi:10.1107/S0365110X63003108.{{cite journal}}: CS1 maint: numeric names: authors list (link)
  31. Weeks, Mary Elvira (1956). The discovery of the elements (6th ed.). Easton, PA: Journal of Chemical Education.
  32. Weeks, Mary Elvira (October 1932). "The discovery of the elements. XVI. The rare earth elements". Journal of Chemical Education. 9 (10): 1751. Bibcode:1932JChEd...9.1751W. doi:10.1021/ed009p1751.
  33. "Ytterbium". Royal Society of Chemistry. 2020. Retrieved 4 January 2020.
  34. "Separaton[sic] of Rare Earth Elements by Charles James". National Historic Chemical Landmarks. American Chemical Society. Retrieved 2014-02-21.
  35. 1 2 Urbain, M.G. (1908). "Un nouvel élément, le lutécium, résultant du dédoublement de l'ytterbium de Marignac". Comptes rendus hebdomadaires des séances de l'Académie des Sciences (in French). 145: 759–762.
  36. Urbain, G. (1909). "Lutetium und Neoytterbium oder Cassiopeium und Aldebaranium – Erwiderung auf den Artikel des Herrn Auer v. Welsbach". Monatshefte für Chemie. 31 (10): 1. doi:10.1007/BF01530262. S2CID   101825980.
  37. von Welsbach, Carl A. (1908). "Die Zerlegung des Ytterbiums in seine Elemente". Monatshefte für Chemie. 29 (2): 181–225. doi:10.1007/BF01558944. S2CID   197766399.
  38. Hedrick, James B. "Rare-Earth Metals" (PDF). USGS. Retrieved 2009-06-06.
  39. Halmshaw, R. (1995). Industrial radiology: theory and practice. Springer. pp. 168–169. ISBN   978-0-412-62780-4.
  40. NIST (2013-08-22) Ytterbium Atomic Clocks Set Record for Stability.
  41. Peik, Ekkehard (2012-03-01). New "pendulum" for the ytterbium clock. ptb.de.
  42. "NIST ytterbium atomic clocks set record for stability". Phys.org. August 22, 2013.
  43. Ostby, Eric (2009). Photonic Whispering-Gallery Resonations in New Environments (PDF) (Thesis). California Institute of Technology . Retrieved 21 December 2012.
  44. Grukh, Dmitrii A.; Bogatyrev, V. A.; Sysolyatin, A. A.; Paramonov, Vladimir M.; Kurkov, Andrei S.; Dianov, Evgenii M. (2004). "Broadband Radiation Source Based on an Ytterbium-Doped Fibre With Fibre-Length-Distributed Pumping". Quantum Electronics. 34 (3): 247. Bibcode:2004QuEle..34..247G. doi:10.1070/QE2004v034n03ABEH002621. S2CID   250788004.
  45. Kouznetsov, D.; Bisson, J.-F.; Takaichi, K.; Ueda, K. (2005). "Single-mode solid-state laser with short wide unstable cavity". Journal of the Optical Society of America B . 22 (8): 1605–1619. Bibcode:2005JOSAB..22.1605K. doi:10.1364/JOSAB.22.001605.
  46. McCumber, D.E. (1964). "Einstein Relations Connecting Broadband Emission and Absorption Spectra". Physical Review B. 136 (4A): 954–957. Bibcode:1964PhRv..136..954M. doi:10.1103/PhysRev.136.A954.
  47. Becker, P.C.; Olson, N.A.; Simpson, J.R. (1999). Erbium-Doped Fiber Amplifiers: Fundamentals and Theory. Academic press.
  48. Kouznetsov, D. (2007). "Comment on Efficient diode-pumped Yb:Gd2SiO5 laser". Applied Physics Letters. 90 (6): 066101. Bibcode:2007ApPhL..90f6101K. doi:10.1063/1.2435309.
  49. Zhao, Guangjun; Su, Liangbi; Xu, Jun; Zeng, Heping (2007). "Response to Comment on Efficient diode-pumped Yb:Gd2SiO5 laser". Applied Physics Letters. 90 (6): 066103. Bibcode:2007ApPhL..90f6103Z. doi: 10.1063/1.2435314 .
  50. Koponen, Joona J.; Söderlund, Mikko J.; Hoffman, Hanna J. & Tammela, Simo K. T. (2006). "Measuring photodarkening from single-mode ytterbium doped silica fibers". Optics Express. 14 (24): 11539–11544. Bibcode:2006OExpr..1411539K. doi: 10.1364/OE.14.011539 . PMID   19529573. S2CID   27830683.
  51. Bisson, J.-F.; Kouznetsov, D.; Ueda, K.; Fredrich-Thornton, S. T.; Petermann, K.; Huber, G. (2007). "Switching of Emissivity and Photoconductivity in Highly Doped Yb3+:Y2O3 and Lu2O3 Ceramics". Applied Physics Letters. 90 (20): 201901. Bibcode:2007ApPhL..90t1901B. doi:10.1063/1.2739318.
  52. Sochinskii, N.V.; Abellan, M.; Rodriguez-Fernandez, J.; Saucedo, E.; Ruiz, C.M.; Bermudez, V. (2007). "Effect of Yb concentration on the resistivity and lifetime of CdTe:Ge:Yb codoped crystals" (PDF). Applied Physics Letters. 91 (20): 202112. Bibcode:2007ApPhL..91t2112S. doi:10.1063/1.2815644. hdl: 10261/46803 .
  53. Samson, Bryce; Carter, Adrian; Tankala, Kanishka (2011). "Doped fibres: Rare-earth fibres power up". Nature Photonics. 5 (8): 466. Bibcode:2011NaPho...5..466S. doi:10.1038/nphoton.2011.170.
  54. "Fiber for Fiber Lasers: Matching Active and Passive Fibers Improves Fiber Laser Performance". Laser Focus World. 2012-01-01.
  55. Olmschenk, S. (Nov 2007). "Manipulation and detection of a trapped Yb171+ hyperfine qubit". Physical Review A. 76 (5): 052314. arXiv: 0708.0657 . Bibcode:2007PhRvA..76e2314O. doi:10.1103/PhysRevA.76.052314. S2CID   49330988.
  56. "Quantinuum | Hardware". www.quantinuum.com. Retrieved 2023-05-21.
  57. "IonQ | Our Trapped Ion Technology". IonQ. Retrieved 2023-05-21.
  58. Hayes, D. (Apr 2010). "Entanglement of Atomic Qubits Using an Optical Frequency Comb". Physical Review Letters. 104 (14): 140501. arXiv: 1001.2127 . Bibcode:2010PhRvL.104n0501H. doi:10.1103/PhysRevLett.104.140501. PMID   20481925. S2CID   14424109.
  59. Gupta, C.K. & Krishnamurthy, Nagaiyar (2004). Extractive metallurgy of rare earths. CRC Press. p. 32. ISBN   978-0-415-33340-5.
  60. Koch, E. C.; Hahma, A. (2012). "Metal-Fluorocarbon Pyrolants. XIV: High Density-High Performance Decoy Flare Compositions Based on Ytterbium/Polytetrafluoroethylene/Viton®". Zeitschrift für Anorganische und Allgemeine Chemie. 638 (5): 721. doi: 10.1002/zaac.201200036 .
  61. Ganesan, M.; Bérubé, C. D.; Gambarotta, S.; Yap, G. P. A. (2002). "Effect of the Alkali-Metal Cation on the Bonding Mode of 2,5-Dimethylpyrrole in Divalent Samarium and Ytterbium Complexes". Organometallics. 21 (8): 1707. doi:10.1021/om0109915.
  62. Gale, T.F. (1975). "The Embryotoxicity of Ytterbium Chloride in Golden Hamsters". Teratology. 11 (3): 289–95. doi:10.1002/tera.1420110308. PMID   807987.
  63. Ivanov, V. G.; Ivanov, G. V. (1985). "High-Temperature Oxidation and Spontaneous Combustion of Rare-Earth Metal Powders". Combustion, Explosion, and Shock Waves. 21 (6): 656. doi:10.1007/BF01463665. S2CID   93281866.
  64. "Material safety data sheet". espi-metals.com. Retrieved 2009-06-06.

Further reading