Group 3 element

Last updated

Group 3 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
IUPAC group number 3
Name by elementscandium group
CAS group number
(US, pattern A-B-A)
IIIB
old IUPAC number
(Europe, pattern A-B)
IIIA

  Period
4
Scandium sublimed dendritic and 1cm3 cube.jpg
Scandium (Sc)
21 Transition metal
5
Yttrium sublimed dendritic and 1cm3 cube.jpg
Yttrium (Y)
39 Transition metal
6
Lutetium sublimed dendritic and 1cm3 cube.jpg
Lutetium (Lu)
71 Lanthanide
7 Lawrencium (Lr)
103 Actinide

Legend

primordial element
synthetic element
Atomic number color:
black=solid

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.

Contents

The chemistry of the group 3 elements is typical for early transition metals: they all essentially have only the group oxidation state of +3 as a major one, and like the preceding main-group metals are quite electropositive and have a less rich coordination chemistry. Due to the effects of the lanthanide contraction, yttrium and lutetium are very similar in properties. Yttrium and lutetium have essentially the chemistry of the heavy lanthanides, but scandium shows several differences due to its small size. This is a similar pattern to those of the early transition metal groups, where the lightest element is distinct from the very similar next two.

All the group 3 elements are rather soft, silvery-white metals, although their hardness increases with atomic number. They quickly tarnish in air and react with water, though their reactivity is masked by the formation of an oxide layer. The first three of them occur naturally, and especially yttrium and lutetium are almost invariably associated with the lanthanides due to their similar chemistry. Lawrencium is strongly radioactive: it does not occur naturally and must be produced by artificial synthesis, but its observed and theoretically predicted properties are consistent with it being a heavier homologue of lutetium. None of the group 3 elements have any biological role.

Historically, sometimes lanthanum (La) and actinium (Ac) were included in the group instead of lutetium and lawrencium, because the electron configurations of many of the rare earths were initially measured wrongly. This version of group 3 is still commonly found in textbooks, but most authors focusing on the subject are against it. Some authors attempt to compromise between the two formats by leaving the spaces below yttrium blank, but this contradicts quantum mechanics as it results in an f-block that is 15 elements wide rather than 14 (the maximum occupancy of an f-subshell).

Composition

Physical, chemical, and electronic evidence overwhelmingly shows that the correct elements in group 3 are scandium, yttrium, lutetium, and lawrencium: [1] [2] [3] [4] [5] [6] [7] this is the classification adopted by most chemists and physicists who have considered the matter. [2] It was supported by IUPAC in a 1988 report [3] and reaffirmed in 2021. [8] Many textbooks however show group 3 as containing scandium, yttrium, lanthanum, and actinium, a format based on historically wrongly measured electron configurations: [4] Lev Landau and Evgeny Lifshitz already considered it to be "incorrect" in 1948, [5] but the issue was brought to a wide debate only in 1982 by William B. Jensen. [4]

The spaces below yttrium are sometimes left blank as a third option, but there is confusion in the literature on whether this format implies that group 3 contains only scandium and yttrium, or if it also contains all the lanthanides and actinides; [8] [9] [10] [11] [12] either way, this format contradicts quantum physics by creating a 15-element-wide f-block when only 14 electrons can fit in an f-subshell. [8] While the 2021 IUPAC report noted that 15-element-wide f-blocks are supported by some practitioners of a specialised branch of relativistic quantum mechanics focusing on the properties of superheavy elements, the project's opinion was that such interest-dependent concerns should not have any bearing on how the periodic table is presented to "the general chemical and scientific community". [8] In fact, relativistic quantum-mechanical calculations of Lu and Lr compounds found no valence f-orbitals in either element. [13] Other authors focusing on superheavy elements since clarified that the "15th entry of the f-block represents the first slot of the d-block which is left vacant to indicate the place of the f-block inserts", which would imply that this form still has Lu and Lr (the 15th entries in question) as d-block elements under Sc and Y. [14] Indeed, when IUPAC publications expand the table to 32 columns, they make this clear and place Lu and Lr under Y. [15] [16]

As noted by the 2021 IUPAC report, Sc-Y-Lu-Lr is the only form that simultaneously allows for the preservation of the sequence of atomic number, avoids splitting the d-block into "two highly uneven portions", and gives the blocks the correct widths quantum mechanics demands (2, 6, 10, and 14). [8] While arguments in favour of Sc-Y-La-Ac can still be found in the literature, many authors consider them to be logically inconsistent. [4] [2] For example, it has been argued that lanthanum and actinium cannot be f-block elements because their atoms have not begun to fill the f-subshells. [17] But the same is true of thorium which is never disputed as an f-block element, [8] [4] and this argument overlooks the problem on the other end: that the f-shells complete filling at ytterbium and nobelium (matching the Sc-Y-Lu-Lr form), not at lutetium and lawrencium (as in Sc-Y-La-Ac). [18] Lanthanum, actinium, and thorium are simply examples of exceptions to the Madelung rule; not only do those exceptions represent a minority of elements (only 20 out of 118), [18] but they have also never been considered as relevant for positioning any other elements on the periodic table. In gaseous atoms, the d-shells complete their filling at copper (3d104s1), palladium (4d105s0), and gold (5d106s1), but it is universally accepted by chemists that these configurations are exceptional and that the d-block really ends in accordance with the Madelung rule at zinc (3d104s2), cadmium (4d105s2), and mercury (5d106s2). [9] The relevant fact for placement is that lanthanum and actinium (like thorium) have valence f-orbitals that can become occupied in chemical environments, whereas lutetium and lawrencium do not: [6] [19] their f-shells are in the core, and cannot be used for chemical reactions. [20] Thus the relationship between yttrium and lanthanum is only a secondary relationship between elements with the same number of valence electrons but different kinds of valence orbitals, such as that between chromium and uranium; whereas the relationship between yttrium and lutetium is primary, sharing both valence electron count and valence orbital type. [19]

History

The discovery of the group 3 elements is inextricably tied to that of the rare earths, with which they are universally associated in nature. In 1787, Swedish part-time chemist Carl Axel Arrhenius found a heavy black rock near the Swedish village of Ytterby, Sweden (part of the Stockholm Archipelago). [21] Thinking that it was an unknown mineral containing the newly discovered element tungsten, [22] he named it ytterbite. [lower-alpha 1] Finnish scientist Johan Gadolin identified a new oxide or "earth" in Arrhenius' sample in 1789, and published his completed analysis in 1794; [23] in 1797, the new oxide was named yttria. [24] In the decades after French scientist Antoine Lavoisier developed the first modern definition of chemical elements, it was believed that earths could be reduced to their elements, meaning that the discovery of a new earth was equivalent to the discovery of the element within, which in this case would have been yttrium. [lower-alpha 2] Until the early 1920s, the chemical symbol "Yt" was used for the element, after which "Y" came into common use. [25] Yttrium metal, albeit impure, was first prepared in 1828 when Friedrich Wöhler heated anhydrous yttrium(III) chloride with potassium to form metallic yttrium and potassium chloride. [26] [27] In fact, Gadolin's yttria proved to be a mixture of many metal oxides, that started the history of the discovery of the rare earths. [24]

In 1869, Russian chemist Dmitri Mendeleev published his periodic table, which had an empty space for an element above yttrium. [28] Mendeleev made several predictions on this hypothetical element, which he called eka-boron. By then, Gadolin's yttria had already been split several times; first by Swedish chemist Carl Gustaf Mosander, who in 1843 had split out two more earths which he called terbia and erbia (splitting the name of Ytterby just as yttria had been split); and then in 1878 when Swiss chemist Jean Charles Galissard de Marignac split terbia and erbia themselves into more earths. Among these was ytterbia (a component of the old erbia), [21] which Swedish chemist Lars Fredrik Nilson successfully split in 1879 to reveal yet another new element. [29] [30] He named it scandium, from the Latin Scandia meaning "Scandinavia". Nilson was apparently unaware of Mendeleev's prediction, but Per Teodor Cleve recognized the correspondence and notified Mendeleev. Chemical experiments on scandium proved that Mendeleev's suggestions were correct; along with discovery and characterization of gallium and germanium this proved the correctness of the whole periodic table and periodic law. [31] Metallic scandium was produced for the first time in 1937 by electrolysis of a eutectic mixture, at 700–800 °C, of potassium, lithium, and scandium chlorides. [32] Scandium exists in the same ores that yttrium had been discovered from, but is much rarer and probably for that reason had eluded discovery. [24]

The remaining component of Marignac's ytterbia also proved to be a composite. In 1907, French scientist Georges Urbain, [33] Austrian mineralogist Baron Carl Auer von Welsbach, and American chemist Charles James [34] all independently discovered a new element within ytterbia. Welsbach proposed the name cassiopeium for his new element (after Cassiopeia), whereas Urbain chose the name lutecium (from Latin Lutetia, for Paris). The dispute on the priority of the discovery is documented in two articles in which Urbain and von Welsbach accuse each other of publishing results influenced by the published research of the other. [35] [36] In 1909, the Commission on Atomic Mass, which was responsible for the attribution of the names for the new elements, granted priority to Urbain and adopting his names as official ones. An obvious problem with this decision was that Urbain was one of the four members of the commission. [37] In 1949, the spelling of element 71 was changed to lutetium. [38] [39] Later work connected with Urbain's attempts to further split his lutecium however revealed that it had only contained traces of the new element 71, and that it was only von Welsbach's cassiopeium that was pure element 71. For this reason many German scientists continued to use the name cassiopeium for the element until the 1950s. Ironically, Charles James, who had modestly stayed out of the argument as to priority, worked on a much larger scale than the others, and undoubtedly possessed the largest supply of lutetium at the time. [40] Lutetium was the last of the stable rare earths to be discovered. Over a century of research had split the original yttrium of Gadolin into yttrium, scandium, lutetium, and seven other new elements. [21]

Lawrencium is the only element of the group that does not occur naturally. It was probably first synthesized by Albert Ghiorso and his team on February 14, 1961, at the Lawrence Radiation Laboratory (now called the Lawrence Berkeley National Laboratory) at the University of California in Berkeley, California, United States. The first atoms of lawrencium were produced by bombarding a three-milligram target consisting of three isotopes of the element californium with boron-10 and boron-11 nuclei from the Heavy Ion Linear Accelerator (HILAC). [41] The nuclide 257103 was originally reported. The team at the University of California suggested the name lawrencium (after Ernest O. Lawrence, the inventor of cyclotron particle accelerator) and the symbol "Lw", [41] for the new element; IUPAC accepted their discovery, but changed the symbol to "Lr". [42] In 1965, nuclear-physics researchers in Dubna, Soviet Union (now Russia) reported 256103, [43] in 1967, they reported that they were not able to confirm American scientists' data on 257103, [44] and proposed the name "rutherfordium" for the new element. [45] The Dubna group criticised the IUPAC approval of the Berkeley group's discovery as having been hasty. [46] In 1971, the Berkeley group did a whole series of experiments aimed at measuring the nuclear decay properties of element 103 isotopes, [47] in which all previous results from Berkeley and Dubna were confirmed, except that the initial 257103 isotope reported at Berkeley in 1961 turned out to have been 258103. [46] In 1992, the IUPAC Trans-fermium Working Group named the nuclear physics teams at Dubna and Berkeley as the co-discoverers of element 103. When IUPAC made the final decision of the naming of the elements beyond 100 in 1997, it decided to keep the name "lawrencium" and symbol "Lr" for element 103 as it had been in use for a long time by that point. The name "rutherfordium" was assigned to the following element 104, which the Berkeley team had proposed it for. [42]

Characteristics

Chemical

Electron configurations of the group 3 elements
ZElementElectron configuration
21Sc, scandium2, 8,  9,  2[Ar]      3d1 4s2
39Y, yttrium2, 8, 18,  9,  2[Kr]      4d1 5s2
71Lu, lutetium2, 8, 18, 32,  9, 2[Xe] 4f14 5d1 6s2
103Lr, lawrencium2, 8, 18, 32, 32, 8, 3[Rn] 5f14 6d0 7s2 7p1

Like other groups, the members of this family show patterns in their electron configurations, especially the outermost shells, resulting in trends in chemical behavior. Due to relativistic effects that become important for high atomic numbers, lawrencium's configuration has an irregular 7p occupancy instead of the expected 6d, [48] [49] but the regular [Rn]5f146d17s2 configuration is low enough in energy that no significant difference from the rest of the group is observed or expected. [50] [51]

Most of the chemistry has been observed only for the first three members of the group; chemical properties of lawrencium are not well-characterized, but what is known and predicted matches its position as a heavier homolog of lutetium. The remaining elements of the group (scandium, yttrium, lutetium) are quite electropositive. They are reactive metals, although this is not obvious due to the formation of a stable oxide layer which prevents further reactions. The metals burn easily to give the oxides, [52] which are white high-melting solids. They are usually oxidized to the +3 oxidation state, in which they form mostly ionic compounds and have a mostly cationic aqueous chemistry. In this way they are similar to the lanthanides, [52] although they lack the involvement of f orbitals that characterises the chemistry of the 4f elements lanthanum through ytterbium. [53] [54] The stable group 3 elements are thus often grouped with the 4f elements as the so-called rare earths. [52]

Typical transition-metal properties are mostly absent from this group, as they are for the heavier elements of groups 4 and 5: there is only one typical oxidation state and the coordination chemistry is not very rich (though high coordination numbers are common due to the large size of the M3+ ions). This said, low-oxidation state compounds may be prepared and some cyclopentadienyl chemistry is known. The chemistries of group 3 elements are thus mostly distinguished by their atomic radii: [52] yttrium and lutetium are very similar, [55] but scandium stands out as the least basic and the best complexing agent, approaching aluminium in some properties. [52] They naturally take their places together with the rare earths in a series of trivalent elements: yttrium acts as a rare earth intermediate between dysprosium and holmium in basicity; lutetium as less basic than the 4f elements and the least basic of the lanthanides; and scandium as a rare earth less basic than even lutetium. [56] Scandium oxide is amphoteric; lutetium oxide is more basic (although it can with difficulty be made to display some acidic properties), and yttrium oxide is more basic still. [57] Salts with strong acids of these metals are soluble, whereas those with weak acids (e.g. fluorides, phosphates, oxalates) are sparingly soluble or insoluble. [52]

Physical

The trends in group 3 follow those of the other early d-block groups and reflect the addition of a filled f-shell into the core in passing from the fifth to the sixth period. For example, scandium and yttrium are both soft metals. But because of the lanthanide contraction, the expected increase in atomic radius from yttrium to lutetium is reversed; lutetium atoms are slightly smaller than yttrium atoms, but are heavier and have a higher nuclear charge. [58] [59] This makes the metal more dense, and also harder because the extraction of the electrons from the atom to form metallic bonding becomes more difficult. All three metals have similar melting and boiling points. [60] Very little is known about lawrencium, but calculations suggest it continues the trend of its lighter congeners toward increasing density. [61] [62]

Scandium, yttrium, and lutetium all crystallize in the hexagonal close-packed structure at room temperature, [63] and lawrencium is expected to do the same. [64] The stable members of the group are known to change structure at high temperature. In comparison with most metals, they are not very good conductors of heat and electricity because of the low number of electrons available for metallic bonding. [63]

Properties of the group 3 elements [65]
NameSc, scandium Y, yttrium Lu, lutetium Lr, lawrencium
Melting point [66] 1814 K, 1541 °C1799 K, 1526 °C1925 K, 1652 °C1900 K, 1627 °C
Boiling point [60] 3109 K, 2836 °C3609 K, 3336 °C3675 K, 3402 °C ?
Density 2.99 g·cm−34.47 g·cm−39.84 g·cm−3 ? 14.4 g·cm−3
Appearancesilver metallicsilver whitesilver gray ?
Atomic radius [59] 162 pm180 pm174 pm ?

Occurrence

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. The abundance of elements in Earth's crust for group 3 is quite low—all the elements in the group are uncommon, the most abundant being yttrium with abundance of approximately 30  parts per million (ppm); the abundance of scandium is 16 ppm, while that of lutetium is about 0.5 ppm. For comparison, the abundance of copper is 50 ppm, that of chromium is 160 ppm, and that of molybdenum is 1.5 ppm. [67]

Scandium is distributed sparsely and occurs in trace amounts in many minerals. [68] Rare minerals from Scandinavia [69] and Madagascar [70] such as gadolinite, euxenite, and thortveitite are the only known concentrated sources of this element, the latter containing up to 45% of scandium in the form of scandium(III) oxide. [69] Yttrium has the same trend in occurrence places; it is found in lunar rock samples collected during the American Apollo Project in a relatively high content as well. [71]

Monazite, the most important lutetium ore Monazit - Mosambik, O-Afrika.jpg
Monazite, the most important lutetium ore

The principal commercially viable ore of lutetium is the rare-earth phosphate mineral monazite, (Ce,La,etc.)PO4, which contains 0.003% of the element. The main mining areas are China, United States, Brazil, India, Sri Lanka and Australia. Pure lutetium metal is one of the rarest and most expensive of the rare-earth metals with the price about US$10,000/kg, or about one-fourth that of gold. [72] [73]

Production

The most available element in group 3 is yttrium, with annual production of 8,900  tonnes in 2010. Yttrium is mostly produced as oxide, by a single country, China (99%). [74] Lutetium and scandium are also mostly obtained as oxides, and their annual production by 2001 was about 10 and 2 tonnes, respectively. [75]

Group 3 elements are mined only as a byproduct from the extraction of other elements. [76] They are not often produced as the pure metals; the production of metallic yttrium is about a few tonnes, and that of scandium is in the order of 10 kg per year; [76] [77] production of lutetium is not calculated, but it is certainly small. The elements, after purification from other rare-earth metals, are isolated as oxides; the oxides are converted to fluorides during reactions with hydrofluoric acid. [78] The resulting fluorides are reduced with alkaline earth metals or alloys of the metals; metallic calcium is used most frequently. [78] For example:

Sc2O3 + 3 HF → 2 ScF3 + 3 H2O
2 ScF3 + 3 Ca → 3 CaF2 + 2 Sc

Biological chemistry

Group 3 metals have low availability to the biosphere. Scandium, yttrium, and lutetium have no documented biological role in living organisms. The high radioactivity of lawrencium would make it highly toxic to living cells, causing radiation poisoning.

Scandium concentrates in the liver and is a threat to it; some of its compounds are possibly carcinogenic, even though in general scandium is not toxic. [79] Scandium is known to have reached the food chain, but in trace amounts only; a typical human takes in less than 0.1 micrograms per day. [79] Once released into the environment, scandium gradually accumulates in soils, which leads to increased concentrations in soil particles, animals and humans. Scandium is mostly dangerous in the working environment, due to the fact that damps and gases can be inhaled with air. This can cause lung embolisms, especially during long-term exposure. The element is known to damage cell membranes of water animals, causing several negative influences on reproduction and on the functions of the nervous system. [79]

Yttrium tends to concentrate in the liver, kidney, spleen, lungs, and bones of humans. [80] There is normally as little as 0.5 milligrams found within the entire human body; human breast milk contains 4 ppm. [81] Yttrium can be found in edible plants in concentrations between 20 ppm and 100 ppm (fresh weight), with cabbage having the largest amount. [81] With up to 700 ppm, the seeds of woody plants have the highest known concentrations. [81]

Lutetium concentrates in bones, and to a lesser extent in the liver and kidneys. [82] Lutetium salts are known to cause metabolism and they occur together with other lanthanide salts in nature; the element is the least abundant in the human body of all lanthanides. [82] Human diets have not been monitored for lutetium content, so it is not known how much the average human takes in, but estimations show the amount is only about several micrograms per year, all coming from tiny amounts taken by plants. Soluble lutetium salts are mildly toxic, but insoluble ones are not. [82]

Notes

  1. Ytterbite was named after the village it was discovered near, plus the -ite ending to indicate it was a mineral.
  2. Earths were given an -a ending and new elements are normally given an -ium ending.

Related Research Articles

<span class="mw-page-title-main">Actinium</span> Chemical element, symbol Ac and atomic number 89

Actinium is a chemical element; it has symbol Ac and atomic number 89. It was first isolated by Friedrich Oskar Giesel in 1902, who gave it the name emanium; the element got its name by being wrongly identified with a substance André-Louis Debierne found in 1899 and called actinium. Actinium gave the name to the actinide series, a set of 15 elements between actinium and lawrencium in the periodic table. Together with polonium, radium, and radon, actinium was one of the first non-primordial radioactive elements to be isolated.

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">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 oxidation state +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.

<span class="mw-page-title-main">Lawrencium</span> Chemical element, symbol Lr and atomic number 103

Lawrencium is a synthetic chemical element; it has symbol Lr and atomic number 103. It is named in honor of Ernest Lawrence, inventor of the cyclotron, a device that was used to discover many artificial radioactive elements. A radioactive metal, lawrencium is the eleventh transuranic element and the last member of the actinide series. Like all elements with atomic number over 100, lawrencium can only be produced in particle accelerators by bombarding lighter elements with charged particles. Fourteen isotopes of lawrencium are currently known; the most stable is 266Lr with half-life 11 hours, but the shorter-lived 260Lr is most commonly used in chemistry because it can be produced on a larger scale.

The lanthanide or lanthanoid series of chemical elements comprises 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">Periodic table</span> Tabular arrangement of the chemical elements ordered by atomic number

The periodic table, also known as the periodic table of the elements, arranges the chemical elements into rows ("periods") and columns ("groups"). It is an icon of chemistry and is widely used in physics and other sciences. It is a depiction of the periodic law, which says that when the elements are arranged in order of their atomic numbers an approximate recurrence of their properties is evident. The table is divided into four roughly rectangular areas called blocks. Elements in the same group tend to show similar chemical characteristics.

In chemistry, a transition metal is a chemical element in the d-block of the periodic table, though the elements of group 12 are sometimes excluded. The lanthanide and actinide elements are called inner transition metals and are sometimes considered to be transition metals as well.

<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.

<span class="mw-page-title-main">Group (periodic table)</span> Column of elements in the periodic table of the chemical elements

In chemistry, a group is a column of elements in the periodic table of the chemical elements. There are 18 numbered groups in the periodic table; the 14 f-block columns, between groups 2 and 3, are not numbered. The elements in a group have similar physical or chemical characteristics of the outermost electron shells of their atoms, because most chemical properties are dominated by the orbital location of the outermost electron.

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">Mischmetal</span> Pyrophoric rare-earth metal alloy

Mischmetal (from German: Mischmetall – "mixed metal") is an alloy of rare-earth elements. It is also called cerium mischmetal, or rare-earth mischmetal. A typical composition includes approximately 55% cerium, 25% lanthanum, and 15~18% neodymium, with traces of other rare earth metals; it contains 95% lanthanides and 5% iron. Its most common use is in the pyrophoric ferrocerium "flint" ignition device of many lighters and torches, although an alloy of only rare-earth elements would be too soft to give good sparks. For this purpose, it is blended with iron oxide and magnesium oxide to form a harder material known as ferrocerium. In chemical formulae it is commonly abbreviated as Mm, e.g. MmNi5.

A period 7 element is one of the chemical elements in the seventh row of the periodic table of the chemical elements. The periodic table is laid out in rows to illustrate recurring (periodic) trends in the chemical behavior of the elements as their atomic number increases: a new row is begun when chemical behavior begins to repeat, meaning that elements with similar behavior fall into the same vertical columns. The seventh period contains 32 elements, tied for the most with period 6, beginning with francium and ending with oganesson, the heaviest element currently discovered. As a rule, period 7 elements fill their 7s shells first, then their 5f, 6d, and 7p shells in that order, but there are exceptions, such as uranium.

<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 4 element</span> Group of chemical elements

Group 4 is the second group of transition metals in the periodic table. It contains the four elements titanium (Ti), zirconium (Zr), hafnium (Hf), and rutherfordium (Rf). The group is also called the titanium group or titanium family after its lightest member.

In nuclear chemistry, the actinide concept proposed that the actinides form a second inner transition series homologous to the lanthanides. Its origins stem from observation of lanthanide-like properties in transuranic elements in contrast to the distinct complex chemistry of previously known actinides. Glenn Theodore Seaborg, one of the researchers who synthesized transuranic elements, proposed the actinide concept in 1944 as an explanation for observed deviations and a hypothesis to guide future experiments. It was accepted shortly thereafter, resulting in the placement of a new actinide series comprising elements 89 (actinium) to 103 (lawrencium) below the lanthanides in Dmitri Mendeleev's periodic table of the elements.

<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 block of the periodic table is a set of elements unified by the atomic orbitals their valence electrons or vacancies lie in. The term seems to have been first used by Charles Janet. Each block is named after its characteristic orbital: s-block, p-block, d-block, f-block and g-block.

William Barry Jensen is an American chemist and chemical historian.

References

  1. Rothbaum, J. O.; Motta, A.*; Kratish, Y.*; Marks, T.J.* Chemodivergent Organolanthanide Catalyzed C-H a-Mono-Borylation of Pyridines. J. Am. Chem. Soc. 2022, 144, 17086-17096: https://pubs.acs.org/doi/10.1021/jacs.2c06844
  2. 1 2 3 Jensen, William B. (2015). "The positions of lanthanum (actinium) and lutetium (lawrencium) in the periodic table: an update". Foundations of Chemistry. 17: 23–31. doi:10.1007/s10698-015-9216-1. S2CID   98624395. Archived from the original on 30 January 2021. Retrieved 28 January 2021.
  3. 1 2 Fluck, E. (1988). "New Notations in the Periodic Table" (PDF). Pure Appl. Chem. 60 (3): 431–436. doi:10.1351/pac198860030431. S2CID   96704008. Archived (PDF) from the original on 25 March 2012. Retrieved 24 March 2012.
  4. 1 2 3 4 5 William B. Jensen (1982). "The Positions of Lanthanum (Actinium) and Lutetium (Lawrencium) in the Periodic Table". J. Chem. Educ. 59 (8): 634–636. Bibcode:1982JChEd..59..634J. doi:10.1021/ed059p634.
  5. 1 2 L. D. Landau, E. M. Lifshitz (1958). Quantum Mechanics: Non-Relativistic Theory. Vol. 3 (1st ed.). Pergamon Press. pp. 256–7.
  6. 1 2 Wittig, Jörg (1973). "The pressure variable in solid state physics: What about 4f-band superconductors?". In H. J. Queisser (ed.). Festkörper Probleme: Plenary Lectures of the Divisions Semiconductor Physics, Surface Physics, Low Temperature Physics, High Polymers, Thermodynamics and Statistical Mechanics, of the German Physical Society, Münster, March 19–24, 1973. Advances in Solid State Physics. Vol. 13. Berlin, Heidelberg: Springer. pp. 375–396. doi:10.1007/BFb0108579. ISBN   978-3-528-08019-8.
  7. Matthias, B. T. (1969). "Systematics of Super Conductivity". In Wallace, P. R. (ed.). Superconductivity. Vol. 1. Gordon and Breach. pp. 225–294. ISBN   9780677138107.
  8. 1 2 3 4 5 6 Scerri, Eric (18 January 2021). "Provisional Report on Discussions on Group 3 of the Periodic Table" (PDF). Chemistry International. 43 (1): 31–34. doi:10.1515/ci-2021-0115. S2CID   231694898. Archived (PDF) from the original on 13 April 2021. Retrieved 9 April 2021.
  9. 1 2 Thyssen, P.; Binnemans, K. (2011). Gschneidner, K. A. Jr.; Bünzli, J-C.G; Vecharsky, Bünzli (eds.). Accommodation of the Rare Earths in the Periodic Table: A Historical Analysis. Vol. 41. Amsterdam: Elsevier. pp. 1–94. doi:10.1016/B978-0-444-53590-0.00001-7. ISBN   978-0-444-53590-0.{{cite book}}: |journal= ignored (help)
  10. Barber, Robert C.; Karol, Paul J; Nakahara, Hiromichi; Vardaci, Emanuele; Vogt, Erich W. (2011). "Discovery of the elements with atomic numbers greater than or equal to 113 (IUPAC Technical Report)". Pure Appl. Chem. 83 (7): 1485. doi: 10.1351/PAC-REP-10-05-01 .
  11. Karol, Paul J.; Barber, Robert C.; Sherrill, Bradley M.; Vardaci, Emanuele; Yamazaki, Toshimitsu (22 December 2015). "Discovery of the elements with atomic numbers Z = 113, 115 and 117 (IUPAC Technical Report)". Pure Appl. Chem. 88 (1–2): 139–153. doi: 10.1515/pac-2015-0502 .
  12. Pyykkö, Pekka (2019). "An essay on periodic tables" (PDF). Pure and Applied Chemistry. 91 (12): 1959–1967. doi:10.1515/pac-2019-0801. S2CID   203944816 . Retrieved 27 November 2022.
  13. Xu, Wen-Hua; Pyykkö, Pekka (8 June 2016). "Is the chemistry of lawrencium peculiar". Phys. Chem. Chem. Phys. 2016 (18): 17351–5. Bibcode:2016PCCP...1817351X. doi:10.1039/c6cp02706g. hdl: 10138/224395 . PMID   27314425. S2CID   31224634 . Retrieved 24 April 2017.
  14. Smits, Odile R.; Düllmann, Christoph E.; Indelicato, Paul; Nazarewicz, Witold; Schwerdtfeger, Peter (2023). "The quest for superheavy elements and the limit of the periodic table". Nature Reviews Physics. doi:10.1038/s42254-023-00668-y. S2CID   266276980.
  15. Leigh, G. Jeffrey (2009). "Periodic Tables and IUPAC". Chemistry International. 31 (1): 4–6. doi:10.1515/ci.2009.31.1.4 . Retrieved 8 January 2024.
  16. Leigh, G. Jeffrey, ed. (1990). Nomenclature of inorganic chemistry : recommendations 1990. Blackwell Scientific Publications. p. 283. ISBN   0-632-02319-8.
  17. Lavelle, Laurence (2008). "Lanthanum (La) and Actinium (Ac) Should Remain in the d-block". Journal of Chemical Education. 85 (11): 1482–1483. Bibcode:2008JChEd..85.1482L. doi: 10.1021/ed085p1482 .
  18. 1 2 Johnson, David (1984). The Periodic Law (PDF). The Royal Society of Chemistry. ISBN   0-85186-428-7.
  19. 1 2 Jensen, William B. (2000). "The Periodic Law and Table" (PDF). Archived from the original (PDF) on 2020-11-10. Retrieved 10 December 2022.
  20. Wulfsberg, Gary (2000). Inorganic Chemistry. University Science Books. p. 26. ISBN   9781891389016.
  21. 1 2 3 van der Krogt, Peter. "39 Yttrium – Elementymology & Elements Multidict". Elements.vanderkrogt.net. Retrieved 2008-08-06.
  22. Emsley 2001, p. 496
  23. Gadolin, Johan (1794). "Undersökning af en svart tung Stenart ifrån Ytterby Stenbrott i Roslagen". Kongl. Vetenskaps Academiens Nya Handlingar (in Swedish). 15: 137–155.
  24. 1 2 3 Greenwood and Earnshaw, p. 944
  25. Coplen, Tyler B.; Peiser, H. S. (1998). "History of the Recommended Atomic-Weight Values from 1882 to 1997: A Comparison of Differences from Current Values to the Estimated Uncertainties of Earlier Values (Technical Report)". Pure Appl. Chem. IUPAC's Inorganic Chemistry Division Commission on Atomic Weights and Isotopic Abundances. 70 (1): 237–257. doi: 10.1351/pac199870010237 . S2CID   96729044.
  26. Heiserman, David L. (1992). "Element 39: Yttrium" . Exploring Chemical Elements and their Compounds. New York: TAB Books. pp. 150–152. ISBN   0-8306-3018-X.
  27. Wöhler, Friedrich (1828). "Über das Beryllium und Yttrium". Annalen der Physik (in German). 89 (8): 577–582. Bibcode:1828AnP....89..577W. doi:10.1002/andp.18280890805.
  28. Ball, Philip (2002). The Ingredients: A Guided Tour of the Elements. Oxford University Press. pp. 100–102. ISBN   0-19-284100-9.
  29. Nilson, Lars Fredrik (1879). "Sur l'ytterbine, terre nouvelle de M. Marignac". Comptes Rendus (in French). 88: 642–647.
  30. Nilson, Lars Fredrik (1879). "Ueber Scandium, ein neues Erdmetall". Berichte der deutschen chemischen Gesellschaft (in German). 12 (1): 554–557. doi:10.1002/cber.187901201157.
  31. Cleve, Per Teodor (1879). "Sur le scandium". Comptes Rendus (in French). 89: 419–422.
  32. Fischer, Werner; Brünger, Karl; Grieneisen, Hans (1937). "Über das metallische Scandium". Zeitschrift für anorganische und allgemeine Chemie (in German). 231 (1–2): 54–62. doi:10.1002/zaac.19372310107.
  33. Urbain, M. G. (1908). "Un nouvel élément, le lutécium, résultant du dédoublement de l'ytterbium de Marignac". Comptes rendus (in French). 145: 759–762.
  34. "Separation of Rare Earth Elements by Charles James". National Historic Chemical Landmarks. American Chemical Society. Retrieved 2014-02-21.
  35. von Welsbach; Carl Auer (1908). "Die Zerlegung des Ytterbiums in seine Elemente". Monatshefte für Chemie (in German). 29 (2): 181–225. doi:10.1007/BF01558944. S2CID   197766399.
  36. Urbain, G. (1909). "Lutetium und Neoytterbium oder Cassiopeium und Aldebaranium – Erwiderung auf den Artikel des Herrn Auer v. Welsbach". Monatshefte für Chemie (in German). 31 (10): I. doi:10.1007/BF01530262. S2CID   101825980.
  37. Clarke, F. W.; Ostwald, W.; Thorpe, T. E.; Urbain, G. (1909). "Bericht des Internationalen Atomgewichts-Ausschusses für 1909". Berichte der Deutschen Chemischen Gesellschaft (in German). 42 (1): 11–17. doi:10.1002/cber.19090420104.
  38. van der Krogt, Peter. "70. Ytterbium – Elementymology & Elements Multidict". Elements.vanderkrogt.net. Retrieved 4 July 2011.
  39. van der Krogt, Peter. "71. Lutetium – Elementymology & Elements Multidict". Elements.vanderkrogt.net. Retrieved 4 July 2011.
  40. Emsley, John (2001). Nature's building blocks: an A-Z guide to the elements. US: Oxford University Press. pp. 240–242. ISBN   0-19-850341-5.
  41. 1 2 Ghiorso, Albert; Sikkeland, T.; Larsh, A. E.; Latimer, R. M. (1961). "New Element, Lawrencium, Atomic Number 103" (PDF). Phys. Rev. Lett. 6 (9): 473. Bibcode:1961PhRvL...6..473G. doi:10.1103/PhysRevLett.6.473.
  42. 1 2 Greenwood, Norman N. (1997). "Recent developments concerning the discovery of elements 101–111". Pure Appl. Chem. 69 (1): 179–184. doi: 10.1351/pac199769010179 .
  43. Donets, E. D.; Shchegolev, V. A.; Ermakov, V. A. (1965). "Synthesis of the isotope of element 103 (lawrencium) with mass number 256". Atomnaya Énergiya (in Russian). 19 (2): 109.
    Translated in Donets, E. D.; Shchegolev, V. A.; Ermakov, V. A. (1965). "Synthesis of the isotope of element 103 (lawrencium) with mass number 256". Soviet Atomic Energy. 19 (2): 109. doi:10.1007/BF01126414. S2CID   97218361.
  44. Flerov, G. N. (1967). "On the nuclear properties of the isotopes 256103 and 257103". Nucl. Phys. A. 106 (2): 476. Bibcode:1967NuPhA.106..476F. doi:10.1016/0375-9474(67)90892-5.
  45. Karpenko, V. (1980). "The Discovery of Supposed New Elements: Two Centuries of Errors". Ambix. 27 (2): 77–102. doi:10.1179/amb.1980.27.2.77.
  46. 1 2 Barber, R. C.; Greenwood, N. N.; Hrynkiewicz, A. Z.; Jeannin, Y. P.; Lefort, M.; Sakai, M.; Ulehla, I.; Wapstra, A. P.; Wilkinson, D. H. (1993). "Discovery of the transfermium elements. Part II: Introduction to discovery profiles. Part III: Discovery profiles of the transfermium elements". Pure and Applied Chemistry. 65 (8): 1757. doi: 10.1351/pac199365081757 . S2CID   195819585. (Note: for Part I see Pure Appl. Chem., Vol. 63, No. 6, pp. 879–886, 1991)
  47. Eskola, Kari; Eskola, Pirkko; Nurmia, Matti; Albert Ghiorso (1971). "Studies of Lawrencium Isotopes with Mass Numbers 255 Through 260". Phys. Rev. C. 4 (2): 632–642. Bibcode:1971PhRvC...4..632E. doi:10.1103/PhysRevC.4.632.
  48. Eliav, E.; Kaldor, U.; Ishikawa, Y. (1995). "Transition energies of ytterbium, lutetium, and lawrencium by the relativistic coupled-cluster method". Phys. Rev. A . 52 (1): 291–296. Bibcode:1995PhRvA..52..291E. doi:10.1103/PhysRevA.52.291. PMID   9912247.
  49. Zou, Yu; Froese, Fischer C. (2002). "Resonance Transition Energies and Oscillator Strengths in Lutetium and Lawrencium". Phys. Rev. Lett. 88 (18): 183001. Bibcode:2001PhRvL..88b3001M. doi:10.1103/PhysRevLett.88.023001. PMID   12005680. S2CID   18391594.
  50. Jensen, W. B. (2015). "Some Comments on the Position of Lawrencium in the Periodic Table" (PDF). Archived from the original (PDF) on 23 December 2015. Retrieved 20 September 2015.
  51. Xu, W-H.; Pyykkö, P. (2016). "Is the chemistry of lawrencium peculiar?" (PDF). Physical Chemistry Chemical Physics. 18 (26): 17351–17355. Bibcode:2016PCCP...1817351X. doi:10.1039/C6CP02706G. hdl: 10138/224395 . PMID   27314425. S2CID   31224634.
  52. 1 2 3 4 5 6 Greenwood and Earnshaw, pp. 964–5
  53. Wang, Fan; Le-Min, Li (2002). "镧系元素 4f 轨道在成键中的作用的理论研究" [Theoretical Study on the Role of Lanthanide 4f Orbitals in Bonding]. Acta Chimica Sinica (in Chinese). 62 (8): 1379–84.
  54. Xu, Wei; Ji, Wen-Xin; Qiu, Yi-Xiang; Schwarz, W. H. Eugen; Wang, Shu-Guang (2013). "On structure and bonding of lanthanoid trifluorides LnF3 (Ln = La to Lu)". Physical Chemistry Chemical Physics. 2013 (15): 7839–47. Bibcode:2013PCCP...15.7839X. doi:10.1039/C3CP50717C. PMID   23598823.
  55. The Heavy Transition Metals, p. 3
  56. Jørgensen, Christian K. (1988). "Influence of rare earths on chemical understanding and classification". Handbook on the Physics and Chemistry of Rare Earths. Vol. 11. pp. 197–292. doi:10.1016/S0168-1273(88)11007-6. ISBN   9780444870803.
  57. Cotton, S. A. (1994). "Scandium, Yttrium and the Lanthanides: Inorganic and Coordination Chemistry". Encyclopedia of Inorganic Chemistry. John Wiley & Sons. ISBN   0-471-93620-0.
  58. Chistyakov, V. M. (1968). "Biron's Secondary Periodicity of the Side d-subgroups of Mendeleev's Short Table". Journal of General Chemistry of the USSR. 38 (2): 213–214. Retrieved 6 January 2024.
  59. 1 2 Dean, John A. (1999). Lange's handbook of chemistry (Fifteenth ed.). McGraw-Hill, Inc. pp. 589–592. ISBN   0-07-016190-9.
  60. 1 2 Barbalace, Kenneth. "Periodic Table of Elements Sorted by Boiling Point". Environmental Chemistry.com. Retrieved 2011-05-18.
  61. Fournier, Jean-Marc (1976). "Bonding and the electronic structure of the actinide metals". Journal of Physics and Chemistry of Solids. 37 (2): 235–244. Bibcode:1976JPCS...37..235F. doi:10.1016/0022-3697(76)90167-0.
  62. Penneman, R. A.; Mann, J. B. (1976). "'Calculation chemistry' of the superheavy elements; comparison with elements of the 7th period". Proceedings of the Moscow Symposium on the Chemistry of Transuranium Elements: 257–263. doi:10.1016/B978-0-08-020638-7.50053-1.
  63. 1 2 Greenwood and Earnshaw, pp. 946–8
  64. Östlin, A.; Vitos, L. (2011). "First-principles calculation of the structural stability of 6d transition metals". Physical Review B. 84 (11): 113104. Bibcode:2011PhRvB..84k3104O. doi:10.1103/PhysRevB.84.113104.
  65. Lide, D. R., ed. (2003). CRC Handbook of Chemistry and Physics (84th ed.). Boca Raton, FL: CRC Press.
  66. Barbalace, Kenneth. "Periodic Table of Elements Sorted by Melting Point". Environmental Chemistry.com. Retrieved 2011-05-18.
  67. Barbalace, Kenneth. "Periodic Table of Elements". Environmental Chemistry.com. Retrieved 2007-04-14.
  68. Bernhard, F. (2001). "Scandium mineralization associated with hydrothermal lazurite-quartz veins in the Lower Austroalpie Grobgneis complex, East Alps, Austria". Mineral Deposits in the Beginning of the 21st Century. Lisse: Balkema. ISBN   90-265-1846-3.
  69. 1 2 Kristiansen, Roy (2003). "Scandium – Mineraler I Norge" (PDF). Stein (in Norwegian): 14–23. Archived from the original (PDF) on October 8, 2010.
  70. von Knorring, O.; Condliffe, E. (1987). "Mineralized pegmatites in Africa". Geological Journal. 22: 253. doi:10.1002/gj.3350220619.
  71. Stwertka, Albert (1998). "Yttrium" . Guide to the Elements (Revised ed.). Oxford University Press. pp.  115–116. ISBN   0-19-508083-1.
  72. Hedrick, James B. "Rare-Earth Metals" (PDF). USGS. Retrieved 2009-06-06.
  73. Castor, Stephen B.; Hedrick, James B. "Rare Earth Elements" (PDF). Retrieved 2009-06-06.
  74. "Mineral Commodity Summaries 2010: Yttrium" (PDF). United States Geological Survey. Retrieved 2011-07-07.
  75. Emsley 2001, p. 241
  76. 1 2 Deschamps, Y. "Scandium" (PDF). mineralinfo.com. Archived from the original (PDF) on February 25, 2009. Retrieved 2008-10-21.
  77. "Mineral Commodity Summaries 2010: Scandium" (PDF). United States Geological Survey. Retrieved 2011-07-07.
  78. 1 2 Holleman, Arnold F.; Wiberg, Egon; Wiberg, Nils (1985). Lehrbuch der Anorganischen Chemie (in German) (91–100 ed.). Walter de Gruyter. pp. 1056–1057. ISBN   3-11-007511-3.
  79. 1 2 3 Lenntech (1998). "Scandium (Sc) — chemical properties of scandium, health effects of scandium, environmental effects of scandium". Lenntech. Retrieved 2011-05-21.
  80. MacDonald, N. S.; Nusbaum, R. E.; Alexander, G. V. (1952). "The Skeletal Deposition of Yttrium" (PDF). Journal of Biological Chemistry. 195 (2): 837–841. doi: 10.1016/S0021-9258(18)55794-X . PMID   14946195.
  81. 1 2 3 Emsley 2001, pp. 495–498
  82. 1 2 3 Emsley 2001, p. 240

Bibliography