rare-earth element

rare-earth element
/rair"errth"/, Chem.
any of a group of closely related metallic elements, comprising the lanthanides, scandium, and yttrium, that are chemically similar by virtue of having the same number of valence electrons. Also called rare-earth metal.
[1955-60]

* * *

Introduction

      any of a large family of chemical elements consisting of the lanthanoids (the 15 elements from lanthanum to lutetium, atomic numbers 57–71) and, because of chemical similarities to the lanthanoids, the elements scandium (atomic number 21) and yttrium (atomic number 39) of group IIIb. They form a series of 17 chemically similar metals, all but one of which occur in nature. Often they are called rare earths, but this is a misnomer because the term earth properly is applied to the oxide of a metal rather than to the element itself. The rare-earth elements are not even particularly rare, though for a long time they were thought to be.

      The 17 rare-earth elements are: scandium (Sc), yttrium (Y), lanthanum (La), cerium (Ce), praseodymium (Pr), neodymium (Nd), promethium (Pm), samarium (Sm), europium (Eu), gadolinium (Gd), terbium (Tb), dysprosium (Dy), holmium (Ho), erbium (Er), thulium (Tm), ytterbium (Yb), and lutetium (Lu).

      Until the mid-20th century, there was not much use for pure rare-earth elements or compounds except cerium and lanthanum; mixtures of the rare earths, however, had found metallurgical and other uses. By the 1970s three of these elements, yttrium, gadolinium, and europium, were being used in the red phosphors for colour television.

 In the periodic (periodic law) table of the elements (see Figure—>), the rare-earth elements comprise three members of Group IIIb and all 14 members of one of two series of elements generally written apart from the main table. This long series is known as the lanthanoid series because it directly follows lanthanum in a different form of the table. The rare-earth elements all have certain common features in the electronic structure of their atoms, which is the fundamental reason for their chemical similarity.

      The aqueous chemistry of all the rare earths is very similar and changes only slightly in progressing along the lanthanoid series. Because of this similarity, it is difficult to separate individual rare earths. In the few cases in which the rare-earth ion can be oxidized or reduced to another oxidation state, however, chemical separations can be carried out readily. Also, artificial mixtures of elements far apart in the series can be separated easily.

      All of these elements exhibit the +3 oxidation state in their compounds, and in the crystal lattices (the regular arrangement of atoms in the solid forms) of such compounds, one rare-earth ion readily replaces another. The rare-earth metals when heated react strongly with nonmetallic elements to form very stable compounds. They are never found as the free metals in the Earth's crust. Pure minerals of individual rare earths do not exist in nature; all their minerals contain mixtures of the rare-earth elements.

      Promethium is never found in the Earth's crust since it has no stable isotopes and is produced only by nuclear reactions; it can, however, be obtained in quantity from the fission products formed in nuclear reactors.

      The chemical properties of scandium differ sufficiently from those of other rare-earth elements for it to have become segregated from them by the action of geological processes. Scandium seldom is associated with the rare earths in minerals.

History
      The early Greeks defined earths as materials that could not be changed further by the sources of heat then available. Until late in the 18th century, this Greek conception remained strong in chemistry, and oxides of metals such as calcium, aluminum, and magnesium were known as earths and were thought to be elements.

      In 1794, Johan Gadolin, a Finnish chemist, while investigating a rare Swedish mineral, discovered a new earth in impure form, which he believed to be a new element and to which he gave the name ytterbia, from Ytterby, the village where the ore was found. The name, however, was soon shortened to yttria. In 1803, from the same mineral, later named gadolinite in Gadolin's honour, another new earth was reported in the literature independently by several chemists. The new earth became known as ceria, from the asteroid Ceres, which had just been discovered (1801). Since yttria and ceria had been discovered in a rare mineral, and they closely resembled other known earths, they were referred to as the rare earths. Not until 1808 did Sir Humphry Davy (Davy, Sir Humphry, Baronet) demonstrate that the earths as a class were not elements themselves but were compounds of oxygen and metallic elements. Later, a number of chemists verified the existence of ceria and yttria in gadolinite and found that these oxides were also present in a wide variety of other rare minerals. The elements of which yttria and ceria were the oxides were then given the names yttrium and cerium, respectively.

      In the period from 1839 to 1843, Carl Gustaf Mosander (Mosander, Carl Gustaf), a Swedish chemist (and student of Berzelius), found that yttria and ceria were not even the oxides of single elements but were, in fact, mixtures. He reported that if the oxides were dissolved in strong acid and the resulting solution subjected to a long series of fractional precipitations as various salts (oxalates, hydroxides, and nitrates), two new elemental substances could be split off from the main component of each oxide. The two new oxides found in ceria he called lanthana and didymia, and the elements contained in them were named lanthanum and didymium. The new elements found (as their oxides) in yttria he called erbium and terbium, and the oxides were referred to as erbia and terbia. Mosander also was the first to obtain the rare-earth metals themselves from their oxides, although only in impure form. Mosander's researches puzzled the scientists of his time. He seemed to be finding a new group of elements of an entirely different type from any known previously. All formed the same classes of compounds with almost the same properties, and the elements could be distinguished from one another—at that time—only by slight differences in the solubilities and molecular weights of the various compounds.

      In the next few years the literature on the rare earths became confused. There was, for example, considerable controversy for a number of years over the existence of didymium. The situation was considerably clarified in 1859 when an instrument called the spectroscope (spectroscopy) was introduced into the study of the rare earths. This instrument indicated the patterns of light emission or absorption characteristic of the elements, and, with it, didymium was shown to have a characteristic absorption spectrum. From then on determination of spectra of various types became one of the most important tools in following the progress of the fractionation of rare earths. Somehow during this period the names used for the various fractions differed from laboratory to laboratory. Around 1860, by general agreement, it was decided to interchange the names of Mosander's earths, erbia and terbia.

      In 1869, when Mendeleyev (Mendeleyev, Dmitry Ivanovich) first proposed the periodic table, he found it necessary to leave a blank at the position now occupied by scandium. He predicted, however, that a new element would be found to fit that blank in the table, and he also predicted certain properties of the element. The discovery of scandium a few years later (1879) and the agreement of its properties with those predicted by Mendeleyev helped to bring about general scientific acceptance of Mendeleyev's periodic table. Interestingly enough, one of the greatest weaknesses in the table was that it provided no logical place for the lanthanoids, a difficulty that was not resolved for some years.

      From 1843 to 1939 chemical fractionation of the mixed rare-earth salts obtained from many minerals was intensively investigated in both Europe and North America. Mosander's didymia was resolved into several oxides—samaria (samarium; 1879), praseodymia (praseodymium; 1885), neodymia (neodymium; 1885), and europia (europium; 1901). His terbia and erbia were resolved into holmia (holmium; 1878), thulia (thulium; 1879), dysprosia (dysprosium; 1886), ytterbia (ytterbium; 1876), and lutetia (lutetium; 1907).

      During this period many of these elements were discovered independently by more than one investigator, but the credit for the discovery was usually given to the man who first separated sufficient quantities of the oxide to determine some of its properties and who published his results first.

      As the scientists carried out their fractionations, they frequently observed changes in colour, apparent molecular weight, and spectra of the substances. Such changes were mainly responsible for the more than 70 claims for the discovery of new rare-earth elements during this period. Many of the observed changes were brought about by the concentration of different impurities, particularly the transition elements, in various fractions of the series. It is now known that such trace impurities in the rare-earth oxides can give rise to such colour changes and that such oxides can be made to fluoresce strongly and exhibit unique spectra.

      Shortly after Auer von Welsbach (Welsbach, Carl Auer, Freiherr von) isolated praseodymia and neodymia in 1885 he invented an illuminating device that bears his name (Welsbach gas mantle), and a little later he produced a practical lighter flint. Both devices depended upon rare-earth elements. Although minerals rich in rare earths had up to that time been thought to be very rare, the demand for rare earths that developed as a result of Auer von Welsbach's inventions resulted in a worldwide search for rare-earth minerals, and it was found that one of them, monazite, existed in extensive deposits. Monazite, a phosphate of several rare-earth elements, was ideal for Auer von Welsbach's purposes, because it contained a high percentage of the element thorium, which was also used for the mantles. These were prepared by impregnating a cloth fabric with a solution of about 90 percent thorium nitrate, 10 percent cerium trinitrate, and minor amounts of other salts. When heated by a gas flame, these salts were converted to their oxides, which, when heated by the flame, gave off an intense white light. Cerium and iron form an alloy that emits sparks when struck. The discovery of this alloy by Auer von Welsbach started the flint industry. In 1913 about 3,300 tons of monazite were refined to produce the thorium and cerium used in gas mantles and the mixed rare-earth metals for flints and related products.

      The British physicist H.G.J. Moseley (Moseley, Henry Gwyn Jeffreys), while studying the X-ray emission spectra of the elements in 1913–14, found a direct relationship between the X-ray frequencies and the atomic numbers (atomic number) of the elements. This relationship made it possible to assign unambiguous atomic numbers to the elements and to verify their locations in the periodic table. In this way, Moseley was able to show clearly that there could be only 14 lanthanoids following lanthanum, starting with cerium and ending at lutetium, and, at that time, all of the rare-earth elements had been discovered except for element 61. Because no stable isotopes (forms of the element with differing mass) of this substance exist in nature, it was not isolated until 1945, when one of its radioactive isotopes was separated from atomic fission products produced in a nuclear reactor. The element was named promethium after the Greek Titan who stole fire from the gods and gave it to mankind.

      As in most fields of science, the present state of knowledge concerning the rare earths is the result of hundreds of scientists publishing thousands of papers and of the individual scientist making his advances based on the work that had been previously published. There were, of course, a number of men whose outstanding contributions changed the direction of the researches, but space does not permit referring to them by name.

Occurrence and abundance
      The rare-earth elements are not rare in nature. They are found in low concentrations widely distributed throughout the Earth's crust and in high concentrations in a considerable number of minerals. In addition, they are also found in many meteorites, on the Moon, and in the Sun. The spectra of many types of stars indicate that the rare-earth elements are much more abundant in these systems than they are in our solar system. Even promethium-147, which has a half-life (time required for one-half the material to undergo radioactive decay) of only a few years, has been observed in certain stars.

      Cerium is reported to be more abundant in the Earth's crust than tin, and yttrium and neodymium more abundant than lead. Even the relatively scarce lutetium is said to be more abundant than mercury or iodine.

      The rare-earth elements are found as mixtures in almost all massive rock formations, in concentrations of from ten to a few hundred parts per million by weight. The fact that these elements have not been separated into minerals containing individual members of the family at any time in the Earth's history—even after eons of repeated melting and resolidifying, mountain formation and erosion, exposure to hot vapour, and immersion in seawater—attests to the great similarity in properties of these elements. Nevertheless, rock formations resulting from some of these geological processes become enriched or depleted in rare earths at one end of the series or the other, so that an analysis of the relative content of the rare-earth elements is never exactly the same, even for similar rocks (rock) taken from different locations. In general, it has been found that the more basic (or alkaline) rocks contain smaller amounts of rare earths than do the more acid rocks, and it is believed that as these molten basic rocks intrude into the more acidic rocks, the rare earths are partially extracted into the more acidic rocks. Also, as this extraction takes place, the rare-earth elements of lower molecular weight (lanthanum, cerium, praseodymium, and neodymium) are taken up to a greater extent than the heavier elements.

      Analytical methods involving activation analysis (production of artificial radioactivity) and mass spectroscopy (mass spectrometry) (separation of atoms on the basis of mass) have made it possible to make accurate measurements of the relative abundances of these elements, even when they are present in extremely small amounts. Such measurements are of great interest to geophysicists because they supply valuable information about the development of geological formations. The cooling of molten rocks and superheated water solutions that have percolated through rock under great pressure frequently produces minerals containing up to 50 percent rare earths. (For uniformity, these percentages are calculated as if the entire rare-earth content of the mineral were present in the form of oxides.) From the presence and composition of such minerals, geochemists can learn a great deal about the conditions, such as temperature and pressure, to which the rock mass was subjected. Similarly, the relative abundance of rare earths in the rocks on the Moon is of great interest because of what it is expected to reveal about how the Moon was formed and whether all or part of the Moon was molten at any time.

       Abundance of the naturally occurring rare-earth elementsThe average content of rare-earth elements found in certain meteorites (chondrites (chondrite)) and in three types of common rocks is listed in the Table (Abundance of the naturally occurring rare-earth elements). Included also is an estimate of the relative abundance of the elements in terms of the overall rank of all known elements and of their concentration in the Earth's crust. It is now generally accepted that the relative values of the rare-earth elements in chondritic (granular) meteorites represent their overall relative abundance in the Cosmos. The elements with even atomic numbers are much more abundant than the odd-numbered elements. Such information, together with the relative abundance of their isotopes, is of critical importance to astrophysicists because it bears on theories of the origin of the universe and the genesis of the chemical elements.

General properties of the group

Electronic structure
      The 14 elements in the lanthanoid series—from cerium through lutetium—are much alike because the differences in their electronic structures chiefly involve the inner 4f electrons, whereas it is the outer s and p (and sometimes d) electrons that are involved in chemical bonding with other atoms and thereby determine the chemical behaviour of the elements. Although lanthanum atoms contain no 4f electrons, they resemble the atoms of the lanthanoid elements closely, and it is not surprising that lanthanum should behave much as the lanthanoids do (the name lanthanoid, in fact, merely means lanthanum-like). Scandium and yttrium are elements in the same vertical file in the periodic table as lanthanum, and their atoms, too, have somewhat the same electronic structure but fewer filled shells, the outermost electrons in scandium being two 4s electrons and one 3d. In the case of yttrium, however, the outermost electrons are 5s and 4d electrons, respectively.

      Because of their general similarity in atomic structure, scandium, yttrium, lanthanum, and the 14 lanthanoids are very similar chemically. This similarity is the reason they are found together in nature and also the reason they are so frequently classed together as the rare-earth elements.

       Some properties of the rare-earth elements (values recommended by Ames Laboratory) part IThe Table (Some properties of the rare-earth elements (values recommended by Ames Laboratory) part I) lists the lowest energy electronic configurations of the rare-earth elements in gaseous and in condensed states, i.e., as the free metal or in a crystalline or solution form as a compound. In a compound the 5d6s2 electrons (the superscript indicating the number of electrons in the subshell) are the valence, or bonding, electrons. These electrons are involved in the chemical bond and are usually paired with the electrons of the anion (the negative ion included with the rare-earth ion in the compound), with the result that they are no longer closely associated with the rare-earth atom. In the case of the metal, these electrons are free to wander throughout the crystal, being able to carry an electrical current and known, therefore, as conducting electrons. In the case of lanthanoids in the gaseous state, the valence electrons are localized at the individual atoms, but in many cases the 5d electron can switch to the 4f subshell, giving rise to a state of different energy and different electronic configuration.

      As the charge on the nucleus increases across the lanthanoid series, it pulls the various subshells (lanthanide contraction), especially the 5s and 5p subshells, closer to the nucleus. As a result, the radii of the lanthanoid ions decrease as the atomic number increases. This effect is known as the lanthanoid contraction.

Physical properties
      The pure rare-earth metals are bright and silvery. A bar of europium will tarnish almost immediately when exposed to air and will be entirely converted to the powdered oxide in a few days. Lanthanum, cerium, praseodymium, and neodymium also corrode readily in air; bars of these metals become encrusted with a thick layer of oxide in several weeks. Metallic yttrium, gadolinium, and lutetium, on the other hand, remain bright and shiny for years.

Effect of impurities
      The properties of the rare-earth metals are frequently quite sensitive to the presence of impurities; for example, the light lanthanoid metals will corrode much more rapidly if small amounts of calcium or magnesium or rare-earth oxides are present in the metal. The melting points and transition temperatures between different crystal forms (allotropic forms) can be changed drastically, frequently by several hundred degrees, when the metals are alloyed with other elements.

      Small amounts of nonmetallic impurities also affect many of the properties of the rare-earth elements. Several thousand parts per million by weight of oxygen and even smaller amounts of nitrogen in the metals make them brittle. The effect of nonmetallic impurities on physical properties is determined by atomic percentages (that is, by the relative numbers of atoms present) and not by weight percentages; thus, in lutetium 300 parts per million by weight of hydrogen is about 5 percent on an atomic basis, whereas 1,000 parts per million of oxygen in lutetium by weight represents only 1 percent oxygen on the atomic scale.

       Some properties of the rare-earth elements (values recommended by Ames Laboratory) part IIn determining properties of the rare-earth metals it is obviously essential to work with well-characterized samples. The amount of each individual impurity present should be accurately known, as well as the previous history of the sample with regard to temperature and work. If the reported values of physical properties are to have much meaning, this information concerning the samples used must be given. Unfortunately, because of a lack of appreciation of the importance of impurities or a lack of proper equipment to adequately characterize the sample, there is a wide variation in the numbers reported in the literature for certain properties. In the Table (Some properties of the rare-earth elements (values recommended by Ames Laboratory) part I) are the values of the properties that—in the author's opinion—are the best values known. Some properties, such as elastic constants, resistivity, and effective magnetic moments, are very sensitive to temperature and show marked anomalies at and in the neighbourhood of crystal or magnetic transformations. Also, some properties depend on the angle at which they are measured with respect to the principal crystal axes in the metal.

Difference in properties across the group
      In fact, the rare-earth metals do not resemble one another as closely as was generally believed in the early part of the 20th century. Physical properties differ as much across the lanthanoid elements as they do for most other series in the periodic table. The melting point of lutetium, for example, is almost twice that of lanthanum, and the vapour pressures of ytterbium and europium at 1,000° C are millions of times greater than those of lanthanum and cerium. Lanthanum is a superconductor (superconductivity) of electricity at 6 K (−267° C) and gadolinium is a stronger ferromagnet at 0 K than iron. The properties of adjacent pairs of lanthanoid elements do, however, differ in a predictable and usually regular manner. This behaviour makes these metals uniquely valuable in studying theories of the metallic state, the formation of alloys, and the existence of intermetallic compounds. Each of the rare-earth metals readily combines with almost any other metallic element, and the resulting alloys exhibit a wide variety of properties: they can be hard or soft, brittle or ductile, and they can have high or low melting points. Some are extremely pyrophoric (ignite spontaneously), whereas others cause a coating to be formed on the surface of metals such as magnesium that protects the alloys from corrosion at elevated temperatures.

Formation of hydrides (hydride) and oxides
      Rare-earth metals absorb hydrogen to form stable alloylike hydrides in which percentages of the compounds MH2 (having the metal atom, M, in the +2 oxidation state) range from zero to 100. These hydrides, brittle and metallic in appearance, have a bluish tinge. After absorption of hydrogen to yield the composition MH2, further absorption occurs, finally yielding hydrides MH3 (in which the metal atom displays the 3+ oxidation state). During the change from MH2 to MH3, the properties become more saltlike. The amount of hydrogen per unit volume in yttrium hydride is considerably greater than that in water or liquid hydrogen, and this hydride does not develop a partial pressure of hydrogen gas equal to one atmosphere until the alloy has been heated to a white heat. Cerium metal, once the oxide surface film has been broken, absorbs hydrogen at room temperature and decomposes water vapour at higher temperatures, absorbing the hydrogen and reacting with the oxygen to form a layer of Ce2O3 on the surface. The oxides, nitrides, and carbides of the rare-earth elements are soluble in the molten metals, as are the elements oxygen, nitrogen, and carbon. The exact form in which the dissolved substances are present is not known, but it is generally believed that the nonmetallic elements are present as interstitial atoms (atoms inserted in spaces left in the crystal structure). These dissolved nonmetallic elements remain in solid solution over a considerable composition range at temperatures near the melting point. As the metal is slowly cooled, however, the solubility decreases, and the dissolved elements precipitate as a second phase, probably as the M2O3, nonstoichiometric nitrides, and carbides. The diffusion rate (rate of movement) for nonmetallic elements in the metal is low below 800° C and becomes progressively lower as the temperature is lowered. The properties of the metals containing these impurities, therefore, are dependent upon the heat treatment to which they have been subjected.

      Rare-earth metals often exhibit anisotropy—differences in properties depending on which direction in the crystal they are measured—and the heat treatment of the sample is important in producing polycrystalline metal. It is possible by certain heat treatments to produce large grains oriented preferentially in a given direction. When properties that depend on crystal direction are measured on such polycrystalline samples, the results have little meaning unless the amount of preferred orientation is known.

Energy states of lanthanoids
      The spark and arc spectra (patterns of emitted light) of the gaseous lanthanoids are extremely complicated. There are literally tens of thousands of frequencies of light emitted by each of the lanthanoids, and it requires very powerful instruments (spectrographs) to resolve them. This complexity arises from the fact that the lanthanoids have an incomplete inner subshell, and the angular moments (spins and orbital motions) of the electrons in this subshell can combine in many ways to give many different energy states (energy state). In the most complicated case, that of gadolinium, there are 3,432 different states. Any of these states can combine with the many states arising from the three valence electrons, and this condition results in an incredible number of excited energy levels. The emitted frequencies represent transitions between any of these states. The situation is further complicated by the fact that the ionization energies (ionization energy) of these elements are extremely low, with the result that in the arc- and spark-light sources there are a great many ionized atoms, and their complicated spectra fall on top of those of the neutral atoms. Many thousands of these lines remain to be identified. A start has been made on this task, however, and a few levels—including the basic ones—have been identified. A thorough understanding of the energy levels of the rare-earth atoms will be of great value in arriving at a complete theory by which all the properties of an atom can be calculated from basic principles. Furthermore, the complete identification of the spectral lines of the rare-earth elements will be of great assistance to astronomers in identifying the many lines observed in stellar spectra that are believed to indicate the presence of rare-earth elements.

      The first ionization energies (the energy required to remove an electron from the neutral atom) of these elements are also difficult to determine accurately because of the complexity of the rare-earth spectra.

Rare-earth ions in solids and liquids
      The sharp bands in spectra of solid rare-earth elements and compounds are much better understood. These bands arise from transitions between different energy states of the 4f subshell, and the position of the bands seems to be little affected by the atoms surrounding the lanthanoid atoms. For this reason, scientists have been able to use these bands for more than a century to determine whether particular rare-earth ions are present in solids or liquids. The fine line structure of the bands, which can be resolved at low temperatures, is sensitive to the environment, and this effect makes such spectra a valuable tool for studying the forces that exist in solids and liquids.

      The 4f electrons also are responsible for the strong magnetism exhibited by the metals and compounds of the lanthanoids. In the incomplete 4f subshell the magnetic effects of the different electrons do not cancel out each other as they do in a completed subshell, and this factor gives rise to the interesting magnetic behaviour of these elements. At higher temperatures, all the lanthanoids except lutetium are paramagnetic (weakly magnetic), and this paramagnetism frequently shows a strong anisotropy. As the temperature is lowered, many of the metals exhibit a point below which they become antiferromagnetic (antiferromagnetism) (i.e., magnetic moments of the ions are aligned but some are opposed to others), and, as the temperatures are lowered still further, many of them go through a series of spin rearrangements, which may or may not be in conformity with the regular crystal lattice. Finally, at still lower temperatures, a number of these elements become ferromagnetic (ferromagnetism) (i.e., strongly magnetic, like iron). Some of the metals have saturation moments (magnetism observed when all the magnetic moments of the ions are aligned) greater than iron, cobalt, or nickel. They also show a strong anisotropy in their magnetic behaviour depending on the crystal direction. Study of the magnetism of rare-earth elements has had great influence on present-day theories of magnetism.

Electrical conductivity
      The rare-earth metals, with the exceptions of cerium, ytterbium, and europium, have three electrons available for carrying electrical current. The space occupied by these electrons apparently represents more than 85 percent of the volume associated with the atom of each metal. Cerium is reported to have an average of 3.1 conducting electrons, presumably as the result of the existence of some of its atoms in a state in which four electrons are free to move through the metal. Pure cerium under high pressure or at low temperature assumes a high-density form in which the four-electron state assumes more importance. Europium and ytterbium are much less dense than the other lanthanoids, and they have only two conducting electrons; the third valence electron has moved to an inside subshell (4f). In europium this electron half fills this subshell, and in ytterbium it completes it, the two configurations 4f7 and 4f14 being particularly stable. The electrical and chemical properties of these two metals therefore resemble those of magnesium, calcium, strontium, and barium (metals with two conducting electrons) more closely than those of the other lanthanoids.

Nuclear properties
      As a group the rare-earth elements are rich in total numbers of isotopes (isotope), averaging about 20 each. The elements with odd atomic numbers have only one, or at most two, stable isotopes, but those with even atomic numbers have from four to seven stable isotopes. Some of the unstable isotopes are feebly radioactive, having extremely long half-lives. The unstable radioactive isotopes can be produced in many ways; e.g., by fission, neutron bombardment, radioactive decay of neighbouring elements, and bombardment of neighbouring elements with charged particles. The lanthanoid isotopes are of particular interest to nuclear scientists because they offer a rich field for testing theories about the nucleus, especially because many of these nuclei are nonspherical, a property that has a decided influence on nuclear stability. When either the protons or neutrons complete a nuclear shell (that is, arrive at certain fixed values), the nucleus is exceptionally stable—the number of protons or neutrons required to complete a shell being called a magic number. One particular magic number—82 for neutrons—occurs in the lanthanoid series.

Production and application

Sources and extraction
      Though numerous minerals rich in rare earths are found in the Earth's crust, many are extremely rare, and many more are found only in small pockets in more massive rocks. Although such minerals are of considerable research interest they are not used commercially. monazite, a mixed phosphate of calcium, thorium, cerium, and various lanthanoids, occurs in extensive deposits and is one of the main sources used commercially to obtain the light rare-earth elements. Monazite contains about 50 percent by weight rare-earth elements, in the approximate proportions 50 percent cerium, 20 percent lanthanum, 20 percent neodymium, 5 percent praseodymium, and lesser amounts of samarium, gadolinium, and yttrium. It also contains small amounts of the heavy rare-earth elements. The actual amounts of each element in the mineral vary considerably, depending on the point of origin of the monazite, because the various metallic elements can substitute for one another in the crystal lattice. The mineral probably formed as small crystals in rocks as they cooled, but as the mountains eroded away and were washed into the sea, the monazite, being denser than most other materials, settled first, while the lighter materials were carried farther out to sea. Apparently as a result of this action, sandbars containing monazite are found along the coasts of Brazil and southwestern India. Concentrated deposits are also found on certain uplands, which are thought to have been the beaches of ancient seas or oceans and which were later uplifted. Such deposits in massive amounts are found in Australia, in South Africa, and in the United States in South Carolina, Florida, and Idaho, as well as in many other locations. The mineral is dredged or scooped up, pulverized if necessary, and concentrated by flotation methods. Sometimes a magnetic-belt separator is used to pull the more magnetic monazite to one side in order to separate it from the nonmagnetic materials. The monazite is then shipped to rare-earth chemical plants.

      The mineral xenotime, a phosphate of yttrium and various lanthanoids, is frequently found associated with monazite and may constitute from 1 to 10 percent of the mixed minerals. Xenotime is similar to monazite except that the metallic atoms are approximately 50 to 60 percent yttrium, and it contains more heavy lanthanoids than light ones. Xenotime is one of the main sources of the heavy rare earths, and it can be separated from monazite by the magnetic-belt process because it is more magnetic than monazite.

      Another important source of light rare earths and europium is the mineral bastnaesite, a fluorocarbonate of lanthanum and cerium, with smaller amounts of neodymium and praseodymium. It is found in extensive deposits in eastern California. It contains almost no heavy rare earths, but there is enough europium (about 0.1 percent) to supply much of the world demand for this element. The rock is broken up by blasting and then is crushed and ground to a fine powder. The bastnaesite is separated from the other materials by the usual flotation methods and is then treated chemically so that it can be separated into europium, lanthanum, and cerium fractions by liquid–liquid extraction methods (see below Liquid–liquid extraction (rare-earth element)).

      The niobium titanate minerals, such as fergusonite, euxenite, samarskite, and blomstrandine, are rich in the heavy rare-earth elements but are not used much commercially. The same is true of such silicates as gadolinite and allanite. Other commercial sources of rare-earth oxides are certain uranium- and apatite-mining operations in which the rare earths are obtained as a by-product even though the rare-earth content of the ores is low.

      Very little scandium is found in rare-earth minerals. Most of the scandium produced commercially is a by-product from uranium processing—the scandium, which may be present in amounts up to five parts per million, being recovered from the uranium solution. There is, however, a rare mineral thortveitite—found in Norway—that contains up to 34 percent scandia, Sc2O3.

Methods of separation and purification
      Generally, the rare-earth elements exist in dilute solution as triply charged ions, M3+ (in which M represents an atom of any rare-earth element). Quite early, however, it was found that a number of the elements could also exist in M4+ or M2+ form—including Ce4+, Sm2+, Eu2+, and Yb2+. If an element could be oxidized (to the +4 state) or reduced (to the +2 state), then it could be removed (separation and purification) readily from the other rare earths. Between the years 1930 and 1935, for example, about two kilograms of extremely pure europium compounds were prepared by a separation process making use of the +2 oxidation state of europium. Although europium is one of the less abundant rare-earth elements, it was one of the first of the heavier rare earths to become generally available.

      Because the ions of the rare-earth elements are surrounded by tightly bound water molecules in aqueous solution, compounds of the rare earths formed from aqueous solutions have properties much alike, and this similarity is particularly true for adjacent elements. The problem is still further complicated by the fact that one rare-earth ion can be substituted readily for another in crystal lattices, with the result that most precipitates consist of crystals of almost the same rare-earth mixture as the solution. Because of this behaviour, chemists of the 19th and early 20th centuries found it necessary to resort to laborious fractionation processes to isolate individual rare-earth elements. At the time, many different processes were used, such as fractional crystallization, fractional precipitation, fractional decomposition, and fractional extraction. All of these consisted of separating the mixed rare earths into two approximately equal fractions, one of which would be enriched in the lighter elements and the other in the heavier elements. Both fractions would then be put back into solution and the process repeated on each of them. Usually the adjacent inner fractions would be recombined before proceeding to the next stage. Gradually, the lighter rare earths were collected in the beakers toward one end of the system, with the heavier elements concentrated at the other end.

      As the quantity of material in the end beakers became small, it was usually customary to combine equivalent fractions from other similar runs. At this point the first series would be split into several independent groups, and a new fractionation process more suited to the elements in each fraction started. Needless to say, the quantity of a relatively pure rare-earth compound obtained from the end beakers was distinctly limited.

      Fractional separation methods, particularly for adjacent heavy rare earths, are extremely slow and tedious. One investigator, for example, reported that he had recrystallized the bromate salt of a thulium fraction 15,000 times and could see little difference between the first and last fractions. It is now known that even the purest fractions he obtained contained some ytterbium and erbium. If the particular lanthanoid elements are far apart in atomic number, however, the task is simplified. It takes only a few partial precipitations, for example, to obtain a lanthanum–cerium–praseodymium fraction completely free of erbium, thulium, ytterbium, and lutetium. The most basic (nonacidic) of the rare-earth elements, lanthanum, is very well situated in this respect because it occurs at one end of the lanthanoid elements, and a few fractionations suffice to separate a lanthanum–cerium fraction from the other rare earths. Since cerium in its +4 oxidation state has distinctly different chemical properties from a typical lanthanoid in the +3 oxidation state, it can be separated from lanthanum easily by ordinary chemical operations. Consequently, pure lanthanum and cerium compounds have been commercially available for many years, and even today several companies find the fractionation process the most economical method for producing compounds of these elements in ton quantities.

      Ion exchange is a method of separation based on differential absorption and elution (washing off) of substances from certain solid supporting materials, often powdered or finely divided materials held in glass tubes. The technique was first used in the rare-earth field during World War II to separate fission products (fission product) obtained from nuclear reactors. In December 1943 a research group at the Oak Ridge (Tennessee) national laboratory announced that they had separated the mixed rare-earth elements from certain fission products by ion exchange on an organic resin into three fractions. The first fraction was shown to have radioactivity associated with yttrium, and the final peak to have cerium activity. The middle peak was thought to be a combination of the neodymium and element-61 activities. The group at Oak Ridge continued to develop the elution technique for separating fission products both with and without carriers (nonradioactive materials added to carry with them the radioactive isotopes). By the end of the war, they had succeeded in developing the processes so that they could separate the individual rare-earth elements of the cerium group (cerium through element 61) and yttrium. The carriers usually consisted of a few milligrams of each of the corresponding natural rare earths. In the meantime, a group at Iowa State University applied the ion-exchange process to the separation of gram quantities of adjacent rare earths and succeeded in separating the difficult pair praseodymium and neodymium in fairly high purities in gram quantities.

Ion-exchange elution (ion-exchange resin)
      Rare-earth-element separation by the ion-exchange elution process is carried out as follows: At the start of the process, the resin is saturated with singly charged cations, such as ammonium ion, NH4+, or hydrogen ion, H+. Next, a solution of mixed rare-earth ions accompanied by strong acid anions is poured onto the top of the column. When the rare-earth ion encounters the cation-containing resin, it replaces three singly charged cations, and these—along with the strong acid anions—will flow through the column in solution and out the bottom. A band of resin saturated with rare-earth ions forms at the top of the ion exchanger and grows in length as more rare-earth solution is added. There is, however, little separation of individual rare-earth ions as this band forms. An eluant solution containing an anion that complexes with the rare-earth ion is then prepared, for example, an ammonium citrate solution of controlled acidity. This solution is then started flowing through the column to elute the rare-earth band down the column and out the bottom. When the main anions present in ammonium citrate in acid solution, HCit2− or H2Cit, encounter rare-earth cations on the resin, complex ions form; these enter the solution phase, and three singly charged ions deposit on the resin in their place. When the rare-earth complexes reach the ammonium or acid resin, in front of the rare-earth band, the rare-earth ions are again deposited, and the band progresses down the column. The formation constants of the individual rare-earth complexes increase slightly with increasing atomic number. Because the various rare-earth ions on the resin are in equilibrium with the rare-earth complexes in solution as they pass over the band, there is a slight enrichment of heavy rare earths at the front of the band. As the band progresses down the column, this enrichment continues. At the same time, the band grows in length, since ammonium and hydrogen ions are also in equilibrium with the resin and their ions will deposit along with the rare-earth ions. After the band has travelled many band lengths, each rare earth exhibits a bell-shaped elution curve (concentration of rare-earth ions versus volume of eluant leaving the column) and these individual rare-earth bands travel down the column at different rates. The bands overlap badly at first, but after travelling many band lengths, they pull completely apart. The area under each curve is, of course, constant, because the amount of each rare earth on the column does not change, but the concentration of the rare earth in the resin gets less and less relative to the ammonium and hydrogen ions on the resin.

      With this type of separation, the original mixed rare-earth band must be quite narrow because the band has to travel many band lengths on a given column. The ions in the eluant are constantly overrunning the bands, with the result that large quantities of solution are needed; and the solution coming out the bottom of the column containing the successive pure rare earths is extremely dilute in rare earths. Such a process is obviously ideal for separating radioactive tracers, which one can count by means of radioactivity, and this process is frequently used in analytical chemistry, where only small amounts of the rare earths are separated. When it is necessary to obtain large amounts of rare earths in high purity, this process is not effective. It has the disadvantage of requiring far more chemicals than the displacement method developed later and described below. Furthermore, this process is not particularly adaptable to being scaled up to produce large quantities of ultrapure rare earths, nor is it well suited for recycling the water and chemicals. It does not give the purity of the individual rare earth that displacement methods can achieve. Finally, the elution process is slow compared with the displacement method.

Ion-exchange band displacement
      The band displacement method of separating individual rare-earth elements was first published in 1952. This process is capable of being scaled up to handle any quantity of rare earths. The mixture can be resolved so that 98 or 99 percent of each individual rare earth can be recovered with less than 0.1 percent of other rare-earth impurities; and, if the rare earths are taken from the middle third of the bands, the sum total of other rare earths can be kept as low as 0.0001 percent. The same resins and type of equipment are used in this process as in the elution technique. Two strong chemical constraints, however, are imposed at the top and bottom edges of the rare-earth band. The eluant contains a strong complexing ion—such as a chelating agent (chelate), an organic molecule that wraps itself around the rare-earth ion, replacing all or most of the adjacent water molecules. The first constraint requires that the formation constants of the rare-earth complexes formed should be large enough so that, when the chelating agent encounters the top edge of the rare-earth band, it complexes in a short distance all of the rare-earth ions, moving them into solution and replacing them on the resin with the cation of the eluant. (The formation constant, however, should not be so large as to remove all the rare-earth ions from the solution phase.) The second chemical constraint occurs at the bottom edge of the rare-earth band: the original resin bed, called the retaining bed, down which the rare-earth band is moving, must have cations on its exchange sites that form a much tighter soluble complex with the chelating ion than do the rare-earth ions. Under this constraint the rare-earth complex promptly breaks up at the point where it encounters the retaining bed, and the rare-earth ions completely deposit in the bed, simultaneously removing an equivalent amount of the retaining-bed cations. With these constraints, the rare-earth band, after spreading out slightly to reach equilibrium, remains of constant length, with sharp top and bottom edges, no matter how far down the column it travels. The elution curve is flat-topped (rare-earth concentration remains constant over almost the entire band when plotted against volume of elute leaving the bottom of the column), and the percentage of rare-earth ions in the rare-earth band on the active sites of the resin is close to 100 percent. Here again, there must be a slight difference in the formation constants of the rare-earth chelates, so that the rare-earth ions are constantly interchanging as the eluant flows by the rare-earth band. As the band moves, the individual rare earths separate into individual flat-topped bands, which ride head to tail and never pull apart. By the time the band has travelled a tenth of its length, most of the heavy rare earths are already to be found in the front segments of the total rare-earth band, and, by the time it has travelled one length, all the individual rare earths are in separate bands, which overlap only slightly to give a narrow region consisting of a binary mixture of adjacent elements. These mixed regions, of course, must be recycled. By having a series of columns, however, the original rare-earth band can be made very long, and, since the overlap regions are independent of band length, the bulk of each successive rare earth comes out the bottom of the column in high purity.

      A number of companies have adopted the displacement process and, using it, have made available highly pure salts of the rare-earth elements of atomic number 59 and above (all the elements from praseodymium through lutetium) in any quantity at reasonable prices. This process has the distinct advantage of allowing the water and the eluting chemicals to be recycled and used over again. One long absorbed band can follow another down a series of columns if a short retaining-bed section is continually regenerated between the absorbed bands.

      The first chelating agent used was ammonium citrate at such a low acidity that the citrate ion, Cit3−, predominates in the solution. At this acidity the complex chelate ion, MCit23−, forms. This process works well, but in 1954 it was improved by using a buffered ammonium solution of ethylenediaminetetraacetic acid (EDTA) with a cupric-ion retaining bed. A number of other chelating agents and types of retaining beds have also been investigated. Many of these work well, but none is markedly superior to EDTA. Some of these other systems, however, are used for certain regions of the rare-earth series because, for these regions, the processing is accomplished more quickly and cheaply.

Liquid–liquid extraction
      Liquid–liquid extraction methods also find important applications in the rare-earth industry. The basic principles involved are similar to those operating in the ion-exchange processes. An organic solvent, such as tributyl phosphate, flows countercurrent (countercurrent distribution) to an aqueous stream containing the mixed rare-earth salts. Rare-earth complexes are formed with formation constants that vary somewhat across the series. The rare-earth ions can complex with their own anions to form neutral molecules that are soluble in the organic phase, or they can complex with molecules of the organic solvent and thereby join the organic stream. If desired, an organic chelating agent can be added to form complexes with the rare-earth ions. These complexes should be soluble in the organic liquid. As the aqueous phase flows past the immiscible organic stream, an equilibrium is set up between the rare-earth ions in the aqueous solution and the complex ions in the organic solvent. As the two streams flow past each other, the heavy rare-earth elements concentrate in one stream and the lighter ones in the other.

      The equilibrium constant for the exchange of one rare-earth ion for another is usually small, with the result that the ions have to exchange with the complex many times before a clear-cut separation between two rare-earth ions is achieved. This process necessitates that the two liquids be in contact with each other through many stages. If the equilibrium constant is equal to 1, no separation will take place, and for adjacent rare earths it is difficult to find complexes—except in special cases—that differ much from that value.

      The liquid–liquid extraction process suffers from the disadvantage that for a given system only one cut is made in the rare-earth series, and if 15 pure rare earths were desired 14 cuts would have to be made. It also suffers from the disadvantage that the distance the rare-earth ions must travel in order to complete one exchange is many times that required in the ion-exchange columns. It has the advantages, however, that more concentrated solutions can be used and that the process is more economical for handling large quantities of materials. So far, it has found application mainly in special cases. It is used in some industries to concentrate the total rare earths where their abundance in the original materials is low. It is also used for separating certain elements, such as lanthanum, cerium, europium, and yttrium, with which favourable equilibrium constants are found. This is the case for cerium and europium, because they can be extracted as Ce4+ and Eu2+, respectively. Yttrium is not a lanthanoid, and its position in the rare-earth series can be changed by using different organic solvents or complexing agents. First, a complex is used that separates the yttrium and heavy lanthanoids from the light lanthanoids, and then a different system is used whereby the yttrium is shifted with respect to the lanthanoid series, so that it can be separated from the heavy lanthanoids.

      The liquid–liquid extraction system has not been successful in separating the adjacent heavy rare earths in the quantities desired. If ultrahigh-purity rare earths are required, it is common practice—even in those cases where liquid–liquid extraction methods have been used—to place the somewhat impure rare earth on an ion-exchange column and to use the displacement method for further purification.

Preparation of pure metals
Early metal- reduction methods
      It is relatively easy to reduce anhydrous halides of the rare earths to metals. What is difficult, however, is to reduce them to high-purity metals in ingot form. The rare-earth metals have a great affinity for the nonmetallic elements—hydrogen, boron, carbon, nitrogen, oxygen, silicon, sulfur, phosphorus, chlorine, and bromine—and form very stable compounds with them. If a small amount of rare-earth metal is added to most other metals containing these elements present as impurities, it reacts with the impurities and removes them by gathering them together in nodules or transferring them to the slag phase. There has been a steady market for misch metal, a mixed rare-earth alloy, since Auer von Welsbach's time. A small addition of this alloy greatly improves the mechanical properties of many impure metals or alloys.

      Also, hot rare-earth turnings (chips or curls from machining) can be used to produce extremely pure helium, neon, and argon by removing hydrogen, oxygen, nitrogen, carbon dioxide, and hydrocarbon vapours. As is often the case with the rare earths, however, other—and cheaper—materials perform this function equally well, and for this reason the rare-earth elements are seldom used for this purpose.

      Finally, molten rare-earth metals dissolve almost all other metals and react with most compounds. They come close to being the hypothetical universal solvent of the ancients. The molten metal attacks any crucible in which it is melted, and the final product generally is a rare-earth-rich alloy of the crucible elements.

      Mosander (Mosander, Carl Gustaf), in 1826, was the first to reduce a rare earth to a metal. He used a metallothermic reaction (heating with active metals) to reduce anhydrous chlorides made from his ceria with metallic sodium or potassium. His yields were low, 26 percent, and the metal existed as small nuggets in a solid slag, from which they could be separated only with difficulty. The metal was very impure; it contained considerable amounts of sodium or potassium and iron and other crucible materials. It also contained considerable amounts of hydrogen, oxygen, nitrogen, and carbon, as well as a mixture of the ceria group of rare earths.

      During the next hundred years, as the individual rare earths were discovered and separated, a number of scientists reduced many of the lighter rare earths to the metallic form using the metallothermic process—but sometimes varying it by substituting calcium, magnesium, and aluminum as the reductants and anhydrous fluoride salts as the reactants. Because of the scarcity of pure rare earths, however, as well as the difficulty in finding suitable crucible materials and the poor equipment for keeping out atmospheric gases, the metals were still so impure that no extensive studies could be made of their properties.

      In 1935, samples of the purest rare-earth chlorides available were reduced to metals at relatively low temperatures in glass capsules with potassium vapour. This process gave free metals in the form of fine powder imbedded in potassium chloride; no attempt was made to separate the metal from the potassium chloride, because only such properties as crystal structure, density, and magnetic susceptibility were under investigation. Potassium chloride acted as an internal standard in the X-ray investigations, and magnetic susceptibilities could be corrected for the potassium chloride present. Although these metals were not really pure by modern standards—they contained appreciable amounts of potassium and rare-earth impurities—they yielded values for the lattice constants and densities of most of the rare-earth metals that lie within 1 percent of the best modern values.

      In 1875, the first successful preparation of rare-earth metals by an electrolytic process (electrolysis) was reported. About five grams each of cerium, lanthanum, and didymium (neodymium and praseodymium) in compact form were prepared by electrolyzing the fused chlorides covered with layers of ammonium chloride. The electrolytic technique was later improved, and, in the period 1902–05, misch metal, cerium, lanthanum, praseodymium, neodymium, and samarium were prepared. In 1906, Auer von Welsbach (Welsbach, Carl Auer, Freiherr von) started the commercial production of lighter flints, for which the misch metal was electrolytically reduced. In the years 1923–26, several improvements in the cell designs were made, and somewhat purer samples of lanthanum, cerium, and neodymium were prepared, along with some yttrium, although most of the latter metal deposited as powder.

      The electrolytic process suffers from much the same difficulties as the metallothermic methods. It is difficult to find electrodes and cell materials that will stand up to molten rare earths and at the same time not introduce impurities into the ingot. It is also difficult to design cells that exclude all the atmospheric elements. The method works best for the low-melting rare earths, with which the cells can be kept sufficiently hot so that a molten pool of the metal forms in the bottom of the cell. With the higher melting rare earths, only powdered metal is formed, and it is difficult to separate it from the electrolyte in a pure form.

      In 1931, a cell especially designed for producing much purer metals and also capable of reducing the halides of the heavier rare earths was employed to produce a quantity of cerium that contained only a small percentage of impurities and, somewhat later, the same apparatus was used to produce a number of other rare-earth metals, including europium, gadolinium, and yttrium.

      By 1939 most of the rare-earth metals had been made in fair purity, and a number of their properties, such as magnetic behaviour, melting point, density, crystal structure, and chemical reactivity had been studied. All of these metals contained small amounts of metallic impurities and unknown amounts of nonmetallic impurities. Most of these impurities were not reported because analytical methods to determine them had not been developed at that time. Almost no work had been done on the properties of the rare-earth alloys except for those of cerium and lanthanum.

Modern techniques for producing ultrapure rare-earth metals
      As purer rare-earth metals are produced, it is increasingly clear that many of their properties are extremely sensitive to small amounts of impurities. This phenomenon is particularly true with regard to magnetic and to nonmetallic impurities. For many industrial uses extreme purity is not required—nor even desired—since less pure metals can be produced much more cheaply. On the other hand, the presence of impurities can be critical in metal produced for research purposes, especially when experimental properties are being compared with predicted values, or in metal to be incorporated into solid-state devices. The processes described below are those used in making research-grade metal. If less pure metal is satisfactory, many of the steps described can be omitted, and the process can be terminated at the point where the desired purity is attained.

      One especially favoured reduction process utilizes metallic calcium (Ca) and the rare-earth fluoride (MF3. The reaction is as in the following equation:

      Other metallothermic processes, however, can also be used, such as lithium (Li) metal and the rare-earth chloride (MCl3):

      Variations of these methods, using lithium, sodium, potassium, or calcium as the reducing agent and any halide of the rare earth for the reactant, also are possible. For these alternative processes to succeed, however, it must be possible to separate the metal from the slag cleanly without introducing impurities, and all sources of contamination must have been eliminated.

      The problem of obtaining sufficient quantities of highly pure individual rare-earth oxides has been solved by the development of the displacement-band method of separating rare earths on ion-exchange columns described above. If the oxides are obtained from the middle third of the pure rare-earth band, the total of other rare-earth impurities in the metals obtained from them does not exceed 10 parts per million. Fractions taken closer to the band edges contain somewhat larger proportions of such impurities. If the same equipment is used to prepare the different raw materials and to make the metals of a number of different rare earths, great care must be taken to prevent cross contamination of the rare earths.

      Contamination from the crucible cannot be eliminated entirely. tungsten and tantalum make the best crucibles: they are little attacked by molten rare-earth metals at temperatures below 1,000° C, and the crucible material introduced into the rare-earth metals at higher temperatures, if harmful, can be removed by special techniques. Both tungsten and tantalum are available commercially in the form of both crucibles and thin sheets: to prepare them for use, these materials are thoroughly cleaned and baked in a high vacuum to remove impurities that may be adsorbed on their surfaces.

      Introduction of impurities from the atmosphere can be largely eliminated by carrying out all operations in an environment of purified helium and by the use of modern high-vacuum ion pumps instead of oil pumps.

      Finally, if ultrapure metals are to be obtained, the raw materials from which they are made must also be ultrapure or their impurities will end up in the rare-earth ingot. Commercial calcium is doubly distilled at low pressure in an atmosphere of pure argon or helium to remove iron and various nonmetals that it contains, and thereafter it is rigorously protected from carbon dioxide and water vapour. Anhydrous rare-earth halides form oxyhalides upon contact with water vapour; therefore, the preparation of an anhydrous fluoride is carried out in two steps. The first is the passage of dry hydrogen fluoride over the powdered oxide, immediately sweeping away any water formed; and the second is the passage of the pure, dry hydrogen fluoride over the molten fluorides. If this is done, the oxygen content of the fluoride can be kept below ten parts per million.

Standard processes for specific groups
      There are enough differences in the properties of the 17 rare-earth metals that the same reduction process does not work equally well for all of them. If metal containing less than 0.01 weight percent of impurities is desired, each element has to be treated somewhat differently. Many of the operations are the same for all reductions, and if the metals are divided into five groups, standardized operations can be applied for all metals in a group. The groups are as follows: Group I consists of those metals that have low melting points and high boiling points—lanthanum, cerium, praseodymium, and neodymium. Group II consists of those metals having high melting points and high boiling points—gadolinium, terbium, scandium, yttrium, and lutetium. Group III consists of those metals having high melting points, low boiling points, and, in addition, an appreciable vapour pressure at the melting point—dysprosium, holmium, erbium, and scandium. Group IV consists of those metals that have low boiling points—samarium, europium, ytterbium, and thulium. Group V, consisting only of promethium, would be included in Group II, except that serious difficulties result from the intense radioactivity of the metal. All operations with it must be carried out by remote controls, and this is usually done at special installations.

      The calcium-reduction process works well for the metals of Groups I, II, III, and V, but not at all well for Group IV. Most of the trifluorides (SmF3, EuF3, YbF3) of this group are reduced only to MF2, even when a large excess of calcium is used; the resulting material resembles, but is not, the metal. Thulium trifluoride is reduced to the metal, but the high vapour pressure of molten metallic thulium causes difficulty.

      The standard procedure for producing the metal by calcium reduction is to load a tantalum container with enough rare-earth fluoride to yield a metal billet weighing about 300 grams. About 10 percent excess calcium is added to drive the reaction to completion. The crucible is then placed in a furnace with a helium atmosphere and heated above the melting point of the rare-earth metal or of the slag—whichever is greater—and held at that temperature until the reaction is complete and the metallic and slag layers have separated because of differences in their densities. After cooling to room temperature, the crucible is taken out of the furnace in the dry box, cut in two at the metal–slag interface, and all slag is knocked off the metal. Usually a bright metal surface can be obtained. The metal ingot, however, contains small amounts of calcium and rare-earth fluoride as impurities. Group I and Group II metals are then put in another tantalum crucible and replaced in the furnace for the boiling-off process. This time a high vacuum is used, and the metal is heated to about 1,400° to 1,500° C and held there for some time, so that any volatile impurities, particularly calcium and rare-earth fluoride, evaporate. At these temperatures considerable amounts (1 to 3 percent) of tantalum dissolve in the molten metal. The furnace temperature is then slowly lowered until it is just above the melting point of the pure metal, at which temperature it is held for a few minutes to allow most of the tantalum to precipitate onto the walls or sink to the bottom of the crucible. (For Group I metals, the solubility of tantalum is about 50 parts per million or less at the melting point.) The ingot is then removed from the cooled furnace in the dry box, and the tantalum crucible and precipitates are machined off. The resulting ingot usually contains less than 0.01 percent total impurities.

      The Group II metals, because of their higher melting points, still contain some tantalum as an impurity when they solidify; usually this tantalum appears as a second phase, showing up as black dots in the metallographic pictures of the metal. It is possible, however, to purify these metals further by distillation from a tantalum still. The still consists of a short tantalum crucible located in a high-vacuum furnace. Affixed to this crucible is an inverted crucible, which is out of the heating zone of the furnace, its upper part being 400° to 500° C cooler because of radiation losses. The rare-earth metal can then be slowly sublimed (sublimation) (changed from a solid into a gas without passing through the liquid state) and resolidified in the inverted crucible. Because volatile impurities usually have different boiling points and heats of sublimation, it is possible (by choosing the right temperature) to sublime the metal in such a manner that the impurities can be separated from the metal. The nonvolatile tantalum remains behind in the still. Finally, the condenser, with its rather porous crystalline mass of rare-earth metal, is removed from the furnace, and the tantalum (or tungsten) is machined off in the dry box. The porous mass is then arc-melted into a billet on a water-cooled copper hearth under an inert atmosphere.

      The Group III elements cannot be held at their melting points for long: because of their volatility a considerable quantity of the metal is lost. The boiling process therefore is omitted, but sublimation or volatilization works well, and—by the right choice of temperature in the still—both the volatile and inert impurities can be eliminated.

      For the Group IV metals, only a distillation process is used. The pure oxide of the rare earth is dissolved in acid and reprecipitated using ultrapure chemicals to remove traces of calcium and magnesium often introduced from the water and chemicals used in the ion-exchange process. The precipitate is again converted to the oxide and placed in the still. Pure metallic lanthanum, cerium, or misch metal, which has been subjected to the boiling-off process to remove volatile impurities, is added in excess.

      The reaction of europium oxide with lanthanum metal (Eu2O3 + 2La → La2O3 + 2Eu↑) takes place when the mixture is heated. Because the vapour pressure of europium is millions of times greater than that of lanthanum or of the oxide of either element, the metal distills away from the oxides, and the reaction goes practically to completion. If ultrahigh-purity metal is desired, a second distillation is performed.

      Thulium poses special problems. Its melting point is so high that the molten metal acquires a considerable amount of impurity from the crucible. On the other hand, it has such a high vapour pressure at the melting point that it is practically impossible to melt it without losing much of the metal. Thulium is never melted, therefore, but is sublimed to the condenser, on which it forms solid crystals but not compact metal. If a solid bar is desired, the porous metal can be pressed into a tantalum tube and reduced to about half its diameter. The tantalum covering can then be machined off and a bar of compact metal obtained.

Major industrial uses
      The properties of the 17 rare-earth elements in the form of their metals, alloys, or compounds—or some combination thereof—are so varied as to make them valuable for many industrial uses (chemical industry). Many other somewhat less costly materials, however, often will perform just as well; and when this is the case, the rare-earth elements are seldom used for these purposes. Only when their properties are unique is the extra cost justified industrially.

Use as catalysts (catalyst)
      Millions of tons of rare earths have been used annually in the United States to produce catalysts for the cracking of crude petroleum. The natural mixture of rare earths obtained from the minerals accounted for about 20 percent of that total, and the remaining 80 percent was made up of special mixtures of lanthanum, praseodymium, neodymium, and samarium. Rare-earth catalysts have been repeatedly recommended for use in numerous organic reactions, including the hydrogenation of ketones to form secondary alcohols, the hydrogenation of olefins to form alkanes, the dehydrogenation of alcohols and butanes, and the formation of polyesters. The extent to which these catalysts are used in industry seldom is made public, but there is no doubt that the rare earths show marked catalytic properties.

Use in the glasses (glass) industry
      Another substantial use of rare-earth oxides is in the glass industry. Cerium oxide has been found to be a more rapid polishing agent for glass than rouge, and several million pounds a year are consumed in the polishing of lenses (lens) for cameras, binoculars, and eyeglasses, as well as in polishing mirrors and television faceplates. Glasses containing lanthanum oxide have very high refractive indexes and low dispersions. Such glasses are used in complex lenses for cameras, binoculars, and military instruments—for the purpose of correcting spherical and chromatic aberrations. Rare-earth oxides often are added to glass melts in order to produce special glasses. Neodymium is added to some glasses to counteract the yellowish tint caused by iron impurities. Very pure neodymium oxide, when added in sufficient quantities (1–5 percent), gives a beautiful purple glass. Praseodymium and neodymium are added to glass to make welders' and glassblowers' goggles, that absorb the bright-yellow light from the sodium flame. The same combination is sometimes added to the glass used in television faceplates to decrease the glare from outside light sources. A beautiful yellow ceramic stain results from the addition of about 3 percent praseodymium oxide to zirconium oxide. Cerium oxide increases the opacity of white porcelain enamels.

Use in the metallurgical industry (metallurgy)
      The metallurgical industry is another heavy user of rare earths. Small amounts of misch metal and cerium have long been added to other metals or alloys (alloy) to remove their nonmetallic impurities. Misch metal added to cast iron makes a more malleable nodular iron. Added to some steels, it makes them less brittle. The addition of misch metal to certain alloys has been reported to increase the tensile strength and improve the hot workability and the high-temperature oxidation resistance. The rare earths are particularly effective in iron–chromium and iron–chromium–nickel alloys to improve a number of their properties, especially their resistance to corrosion and oxidation. Yttrium metal is said to work even better than misch metal in removing impurities from certain materials. The flints of cigarette lighters are an alloy of misch metal and iron.

      The addition of misch metal or pure rare-earth elements to magnesium increases its high-temperature strength and its creep resistance—that is, resistance to slow deformation under prolonged use. This alloy also makes better castings if small amounts of zirconium or other metals are added, and such alloys are used in jet-engine and precision castings. The addition of small amounts of rare-earth elements to aluminum has also been reported to give better castings.

Use in the television industry
      By far the heaviest user of ultrapure separated rare earths is the television industry. It has been found that if a small amount of europium oxide (Eu2O3) is added to yttrium oxide (Y2O3), it gives a brilliant-red phosphor. Colour television screens utilize red, green, and blue phosphors. In the past, a zinc–cadmium sulfide was used as the red phosphor, but it was not completely satisfactory because its fluorescent band was too wide, and it could not be made to fluoresce as intensely as the other phosphors. The Y2O3–Eu2O3 phosphor corrected these disadvantages and made possible much brighter and more natural coloured pictures. This use has been growing in many countries. Many of the early rare-earth screens used europium–yttrium orthovanadate phosphor, but the industry is shifting heavily toward the oxide phosphor. Some television companies have substituted gadolinium oxide for the yttrium oxide. The rare-earth phosphors are also finding use in mercury-arc lights (electric discharge lamp), which are used for sporting events and special street lighting. Instead of the unhealthy-looking blue light of the mercury arc, the phosphors give an intense white radiation similar to daylight. Considerable amounts of mixed rare-earth fluorides are used to make cored carbon rods, which are used as arcs in searchlights and in some of the lights used by the motion-picture industry.

Other applications
      Yttrium-iron garnets are synthetic high-melting silicates that can be fabricated into special shapes for use as microwave filters in the communications industry. Yttrium-aluminum garnets also are being produced at an increasing rate for use both in electronics and as gemstones. Both of these synthetic minerals have much use in the jewelry business. These garnets have a high refractive index and a hardness approaching that of diamond. In a solid crystal form they are amazingly transparent, and they are being cut into imitation diamonds.

      Another significant industrial application of rare earths is in the manufacture of strong permanent magnets. Alloys of cobalt with rare earths, such as cobalt–samarium, produce permanent magnets that are far superior to most of the varieties now on the market. Another relatively recent development is the use of a barium phosphate–europium phosphor in a sensitive X-ray film that forms satisfactory images with only half the exposure.

      Europium, gadolinium, and dysprosium have large capture cross sections for thermal neutrons (neutron capture)—that is, they absorb large numbers of neutrons per unit of area exposed. These elements, therefore, are incorporated into control rods used to regulate the operation of nuclear reactors (nuclear reactor) or to shut them down should they get out of control. In addition, rare-earth elements are used as burnable neutron absorbers to keep the reactivity of the reactor more nearly constant. As uranium undergoes fission, it produces some fission products that absorb neutrons and tend to slow down the nuclear reaction. If the right amounts of rare-earth elements are present, they burn out at about the same rate that other absorbers are formed.

      Yttrium dihydride is used as a moderator in reactors to slow down neutrons. Certain rare earths are also used in shielding materials because of their high nuclear cross sections. Scandium metal is used as a neutron filter that allows neutrons only of a certain energy (two kiloelectron volts) to pass through.

      Complexes of europium, praseodymium, or ytterbium with derivatives of camphor are useful reagents for analysis of optically-active organic compounds, which often are obtained as mixtures containing unknown proportions of two components that differ only in that their molecular structures are mirror images of each other. Determination of these proportions can be very difficult, but the rare-earth complexes provide asymmetric environments in which each component absorbs electromagnetic radiation of a particular frequency in the presence of a strong magnetic field. Proportions then can be determined by measuring the intensities of the separate absorptions.

      The rare earths have low toxicities and can be handled safely with ordinary care. Solutions injected into the peritoneum will cause hyperglycemia (excess of sugar in the blood), decreased blood pressure, spleen degeneration, and fatty liver. If solutions are injected into muscle about 75 percent of the rare-earth element remains at the site, the remainder going to the liver and skeleton. When taken orally, only a small percentage of a rare-earth element is absorbed into the body. Organically complexed ions are somewhat more toxic than solids or inorganic solutions. As is true for most chemicals, dust and vapours should not be inhaled, nor should they be ingested. Solutions splashed into the eyes should be washed out, and splinters of metal should be removed.

      When handling rare-earth ores or minerals, dust should be avoided because many minerals contain other toxic elements, such as beryllium, thorium, and uranium. Finely divided rare-earth metals can ignite spontaneously, somewhat as magnesium does.

Frank Harold Spedding Ed.

Additional Reading
F.H. Spedding and A.H. Daane (eds.), The Rare Earths (1961), broadly reviews rare-earth chemistry and metallurgy to 1960. Karl A. Gschneider, Jr., Rare Earth Alloys (1961), also provides a review to 1960. C. Michael Lederer and Virginia S. Shirley (eds.), Table of Isotopes, 7th ed. (1978), includes a review of radioactivity of the rare-earth elements. Specific aspects are treated in detail in R.J. Elliott (ed.), Magnetic Properties of Rare Earth Metals (1972); P. Henderson (ed.), Rare Earth Element Geochemistry (1984); Bruce R. Lipin and G.A. McKay (eds.), Geochemistry and Mineralogy of Rare Earth Elements (1989); E.J. Wheelwright (ed.), Promethium Technology (1973); Norman M. Edelstein (ed.), Lanthanide and Actinide Chemistry and Spectroscopy (1980); Karl A. Gschneider, Jr., and LeRoy Eyring (eds.), Handbook on the Physics and Chemistry of Rare Earths (1978); Karl A. Gschneider, Jr. (ed.), Industrial Applications of Rare Earth Elements (1981); and C.K. Gupta and T.S. Krishnan, Rare Earths: Applications and Technology (1988).Frank Harold Spedding Ed.

* * *


Universalium. 2010.

Игры ⚽ Нужно сделать НИР?

Look at other dictionaries:

  • Rare earth element — Rare earth ore, shown with a United States penny for size comparison These rare …   Wikipedia

  • rare earth element — rare earth UK [reə(r) ˈɜː(r)θ] / US [reər ˈɜrθ] or rare earth element UK / US noun [countable] Word forms rare earth : singular rare earth plural rare earths chemistry any of a group of metallic elements that have similar chemical properties and… …   English dictionary

  • rare earth element — n any of a series of metallic elements of which the oxides are classed as rare earths and which include the elements of the lanthanide series and sometimes yttrium and scandium called also rare earth metal …   Medical dictionary

  • rare-earth element — n. any of the series of metallic chemical elements, with consecutive atomic numbers of 57 (lanthanum) through 71 (lutetium) inclusive: see LANTHANIDE: often called rare earth metal …   English World dictionary

  • rare-earth element — noun any element of the lanthanide series (atomic numbers 57 through 71) • Syn: ↑rare earth, ↑lanthanoid, ↑lanthanide, ↑lanthanon • Hypernyms: ↑group, ↑grouping • Hyponyms: ↑ …   Useful english dictionary

  • rare earth element — retųjų žemių elementas statusas T sritis fizika atitikmenys: angl. rare earth element vok. Seltenerdelement, n rus. редкоземельный элемент, m pranc. élément de terres rares, m …   Fizikos terminų žodynas

  • rare-earth element — rare′ earth′ el ement n. chem. any of a group of closely related metallic elements, comprising the lanthanides, scandium, and yttrium, that are chemically similar in having the same number of valence electrons. Also called rare′ earth′ met al •… …   From formal English to slang

  • rare earth element — noun Date: 1924 any of a series of metallic elements of which the oxides are classed as rare earths and which include the elements of the lanthanide series and sometimes yttrium and scandium called also rare earth metal …   New Collegiate Dictionary

  • rare earth element — noun Any of a group of soft, lustrous, sometimes magnetic metals that form trivalent salts; they include the lanthanides plus yttrium, scandium, thorium and sometimes zirconium. Syn …   Wiktionary

  • rare-earth element — /rɛər ɜθ ˈɛləmənt/ (say rairr erth eluhmuhnt) noun 1. one of a group of closely related metallic elements of atomic number 57 to 72 inclusive, often divided into three groups: cerium metals (lanthanum, cerium, praseodymium, neodymium, promethium …  

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”