transition element

transition element
any element in any of the series of elements with atomic numbers 21-29, 39-47, 57-79, and 89-107, that in a given inner orbital has less than a full quota of electrons. Also called transition metal.
[1920-25]

* * *

Any chemical element with valence electrons in two shells instead of only one.

This structure gives them their outstanding ability to form ions containing more than one atom (complex ions, or coordination compounds), with a central atom or ion (often of a transition metal) surrounded by ligands in a regular arrangement. Theories on the bonding in these ions are still being refined. The elements in the periodic table from scandium to copper (atomic numbers 21–29), yttrium to silver (39–47), and lanthanum to gold (57–79, including the lanthanide series) are frequently designated the three main transition series. (Those in the actinide series and beyond, 89–111, also qualify.) All are metals, many of major economic or industrial importance (e.g., iron, gold, nickel, titanium). Most are dense, hard, and brittle, conduct heat and electricity well, have high melting points, and form alloys with each other and other metals. Their electronic structure lets them form compounds at various valences. Many of these compounds are coloured and paramagnetic (see paramagnetism) and (as do the metals themselves) often act as catalysts. See also rare earth metal.

* * *

Introduction
 any of various chemical elements that have valence electrons—i.e., electrons that can participate in the formation of chemical bonds—in two shells instead of only one. While the term transition has no particular chemical significance, it is a convenient name by which to distinguish the similarity of the atomic (atom) structures and resulting properties of the elements so designated. They occupy the middle portions of the long periods of the periodic table (periodic law) of elements (see Figure—>) between the groups on the left-hand side and the groups on the right. Specifically, they form Groups 3 (IIIb) through 12 (IIb).

General properties of the group
      The most striking similarities shared by the 24 elements in question are that they are all metals (metal) and that most of them are hard, strong, and lustrous, have high melting and boiling points, and are good conductors of heat and electricity. The range in these properties is considerable; therefore the statements are comparative with the general properties of all the other elements.

      Many of the elements are technologically important: titanium, iron, nickel, and copper, for example, are used structurally and in electrical technology. Second, the transition elements form many useful alloys (alloy), with one another and with other metallic elements. Third, most of these elements dissolve in mineral acids, although a few, such as platinum, silver, and gold, are called “noble”—that is, are unaffected by simple (nonoxidizing) acids.

      Without exception, the elements of the main transition series (i.e., excluding the lanthanoids and actinoids as specified below) form stable compounds in two or more formal oxidation states.

      The transition elements may be subdivided according to the electronic structures of their atoms into three main transition series, called the first, second, and third transition series, and two inner transition series, called the lanthanoids and the actinoids.

       Some properties of the transition elementsThe first main transition series begins with either scandium (symbol Sc, atomic number 21) or titanium (symbol Ti, atomic number 22) and ends with zinc (symbol Zn, atomic number 30). The second series includes the elements yttrium (symbol Y, atomic number 39) to cadmium (symbol Cd, atomic number 48). The third series extends from lanthanum (symbol La, atomic number 57) to mercury (symbol Hg, atomic number 80). These three main transition series are included in the set of 30 elements often called the d-block transition elements. Because scandium, yttrium, and lanthanum actually do not form compounds analogous to those of the other transition elements and because their chemistry is quite homologous to that of the lanthanoids, they are excluded from the present discussion of the main transition elements. Similarly, because zinc, cadmium, and mercury exhibit few of the properties characteristic of the other transition elements, they are treated separately (see zinc group element). The remaining d-block transition elements and some of their characteristic properties are listed in the Table (Some properties of the transition elements).

      The first of the inner transition series includes the elements from cerium (symbol Ce, atomic number 58) to lutetium (symbol Lu, atomic number 71). These elements are called the lanthanoids (or lanthanides) because the chemistry of each closely resembles that of lanthanum. Lanthanum itself is often regarded as one of the lanthanoids. The actinoid series consists of 15 elements from actinium (symbol Ac, atomic number 89) to lawrencium (symbol Lr, atomic number 103). These inner transition series are covered under rare-earth element and actinoid element (actinide element). For elements 104 and higher, see transuranium element.

      The relative locations (electronic configuration) of the transition elements in the periodic table and their chemical and physical properties can best be understood by considering their electronic structures and the way in which those structures vary as atomic numbers increase.

Atomic orbitals (orbital) of the hydrogen atom
      As noted earlier, the electrons associated with an atomic nucleus are localized, or concentrated, in various specific regions of space called atomic orbitals, each of which is characterized by a set of symbols (quantum numbers) that specify the volume, the shape, and orientation in space relative to other orbitals. An orbital may accommodate no more than two electrons. The energy involved in the interaction of an electron with the nucleus is determined by the orbital that it occupies, and the electrons in an atom distribute themselves among the orbitals in such a way that the total energy is minimum. Thus, by electronic structure, or configuration, of an atom is meant the way in which the electrons surrounding the nucleus occupy the various atomic orbitals available to them. The simplest configuration is the set of one-electron orbitals of the hydrogen atom. The orbitals can be classified, first, by principal quantum number, and the orbitals have increasing energy as the principal quantum number increases from 1 to 2, 3, 4, etc. (The sets of orbitals defined by the principal quantum numbers 1, 2, 3, 4, etc., are often referred to as shells designated K, L, M, N, etc.) For principal quantum number 1 there is but a single type of orbital, called an s orbital. As the principal quantum number increases, there are an increasing number of different types of orbitals, or subshells, corresponding to each: s, p, d, f, g, etc. Moreover, the additional orbital types each come in larger sets. Thus, there is but one s orbital for each principal quantum number, but there are three orbitals in the set designated p, five in each set designated d, and so on. For the hydrogen atom, the energy is fully determined by which orbital the single electron occupies. It is especially notable that the energy of the hydrogen atom is determined solely by the principal quantum number of the orbital occupied by the electron (except for some small effects that are not of concern here); that is, in hydrogen, the electron configurations of the third shell, for example, are equi-energic (of the same energy, whichever one the electron occupies), which is not the case with any of the other atoms, all of which contain two or more electrons.

Atomic orbitals of multi-electron atoms
      To understand the electron configurations of other atoms, it is customary to employ the Aufbau (Aufbau principle) (German: “building up”) principle, the basis of which is that, to achieve a multi-electron configuration, the required number of electrons must be added to the orbitals one at a time, filling the most stable orbitals first, until the total number has been added. Thus, in “building up” the periodic table, one progresses from one element to the next by adding one proton to the nucleus and one electron to the atomic region outside the nucleus. There is one restriction upon this conceptualization, namely, the Pauli exclusion principle, which states that only two electrons may occupy each orbital. Thus there can be no more than two electrons in any s orbital, six electrons in any set of p orbitals, ten electrons in any set of d orbitals, etc. In carrying out this process, however, one cannot simply use the ordering of electron orbitals that is appropriate to the hydrogen atom. As electrons are added they interact with each other as well as with the nucleus, and as a result the presence of electrons in some orbital causes the energy of an electron entering another orbital to be different from what it would be if this electron were present alone. The overall result of these interelectronic interactions (sometimes referred to as shielding) is that the relative order of the various atomic orbitals is different in many-electron atoms from that in the hydrogen atom; in fact, it changes continuously as the number of electrons increases.

      As multi-electronic atoms are built up, the various subshells s, p, d, f, g, etc. of a principal quantum number cease to be equi-energic; they all drop, although not by equal amounts, to lower energies. Overall lowering of energy occurs because the shielding from the nuclear charge that an electron in a particular orbital is given by all of the other electrons in the atom is not sufficient to prevent a steady increase in the effect that the charge in the nucleus has on that electron as the atomic number increases. In other words, each electron is imperfectly shielded from the nuclear charge by the other electrons. In addition the different types of orbitals in each principal shell, because of their different spatial distributions, are shielded to different degrees by the core of electrons beneath them; accordingly, although all of them decrease in energy, they decrease by different amounts, and thus their relative order in energy continuously changes. In order to specify the electron configuration of a particular atom, it is necessary to use the order of orbitals appropriate to the specific value of the atomic number of that atom. The behaviour of the various d and f orbitals is to be especially noted in regard to where the transition elements occur in the periodic table.

      The argon atom (atomic number 18) has an electron configuration 1s2 2s2 2p6 3s2 3p6 (i.e., it has two electrons in the s orbital of the first shell; two in the s and six in the p orbitals of the second shell; two in the s and six in the p orbitals of the third shell: this expression often is abbreviated [Ar] especially in specifying the configurations of elements between argon and krypton, because it represents a common part of the configurations of all these elements). The 3d orbitals are more shielded from the nuclear charge than is the 4s orbital, and, consequently, the latter orbital has lower energy. The next electrons to be added enter the 4s orbital in preference to the 3d or 4p orbitals. The two elements following argon in the periodic table are potassium, with a single 4s electron, and calcium, with two 4s electrons. Because of the presence of the 4s electrons, the 3d orbitals are less shielded than the 4p orbitals; therefore, the first regular transition series begins at this point with the element scandium, which has the electron configuration [Ar] 4s23d1. Through the next nine elements, in increasing order of atomic number, electrons are added to the 3d orbitals until, at the element zinc, they are entirely filled and the electron configuration is [Ar] 3d104s2. The 4p orbitals are then the ones of lowest energy, and they become filled through the next six elements, the sixth of which is the next noble gas, krypton, with the electron configuration 1s2 2s2 2p6 3s2 3p6 4s2 3d10 4 p6, or [Kr].

      Throughout the next period the pattern of variation of the orbital energies is similar to that immediately preceding. When the configuration of the noble gas, krypton, has been achieved, the 5s orbital is more stable than the 4d orbitals. The next two electrons therefore enter the 5s orbital, but then the 4d orbitals fall to lower energy than the 5p orbitals, and the second regular transition series commences with the element yttrium. Electrons continue to be added to the 4d orbitals until those orbitals are entirely filled at the position of the element cadmium, which has an electron configuration [Kr] 4d105s2. The next six electrons enter the 5p orbitals until another noble gas configuration is attained at the element xenon. Analogously to the two preceding periods, the next two electrons are added to the next available orbital, namely, the 6s orbital, producing the next two elements, cesium and barium. At this point, however, the ordering of orbitals becomes more complex than it previously had been, because there are now unfilled 4f orbitals as well as the 5d orbitals, and the two sets have approximately the same energy. In the next element, lanthanum (atomic number 57), an electron is added to the 5d orbitals, but the immediately following element, cerium (atomic number 58), has two electrons in the 4f orbitals and none in the 5d orbitals. Through the next 12 elements the additional electrons enter the 4f orbitals, although the 5d orbitals are of only slightly higher energy. This set of elements, spanning the range from lanthanum, where the 4f orbitals were still vacant or about to be filled, through lutetium, in which the 4f orbitals are completely filled by 14 electrons, makes up the lanthanoids, mentioned above.

      At this point the next available orbitals are the 5d orbitals, and the elements hafnium through gold, the third regular transition series, correspond to the successive filling of these 5d orbitals. Following this series there are again p orbitals (6p) to be filled, and when this is accomplished the noble gas radon is reached.

Molecular (molecule) orbitals
      If two atoms are close together, some of their orbitals may overlap and participate in the formation of molecular orbitals. Electrons that occupy a molecular orbital interact with the nuclei of both atoms: if this interaction results in a total energy less than that of the separated atoms, as is the case if the orbital lies mainly in the region between the two nuclei, the orbital is said to be a bonding orbital and its occupancy by electrons constitutes a covalent bond that links the atoms together in compound formation and in which the electrons are said to be shared. If the occupation of an orbital by electrons raises the energy of the system, as is the case if the orbital lies mainly outside the region between the two nuclei, that orbital is said to be antibonding; the presence of electrons in such orbitals tends to offset the attractive force derived from the bonding electrons.

      A bonding or an antibonding molecular orbital may be disposed along the line passing through the two nuclei, in which case it is designated by the Greek letter σ (sigma); or it may occupy regions approximately parallel to that line and be designated π (pi).

Discovery of the transition elements
      The most abundant transition element in the Earth's solid crust is iron, which is fourth among all elements and second (to aluminum) among metals in crustal abundance. The elements titanium, manganese, zirconium, vanadium, and chromium also have abundances in excess of 100 grams (3.5 ounces) per ton. Some of the most important and useful transition elements have very low crustal abundances—e.g., tungsten, platinum, gold, and silver.

      Four of the regular transition elements were known to the ancients: iron (ferrum), copper (cuprum), silver (argentum), and gold (aurum). Their chemical symbols (Fe, Cu, Ag, Au), in fact, are derived from their alchemical (Latin) names rather than their contemporary names. The other regular transition elements were discovered (or recognized as elements) after the early 18th century. The transition element most recently discovered in nature is rhenium (atomic number 75), which in 1925 was detected in platinum ores and in the niobium mineral columbite.

      Technetium (atomic number 43) is the only d-block element that has not been isolated from the Earth's crust. All isotopes of technetium are radioactive; the half-life of even the stablest isotope, technetium-97, is too short to permit the survival of primordial technetium in the Earth's crust, and claims that it has been isolated or detected there must be considered erroneous. Technetium can be isolated in considerable quantities, however, from the fission products of nuclear reactors, and it is at least as readily available for chemical study as the naturally occurring similar element rhenium, of which there are no concentrated ores.

Transition-metal catalysts
      One important use of transition metals and their compounds is as catalysts (catalysis) for a variety of industrial processes, mostly in the petroleum and polymer (plastics, fibres) industries, in which organic molecules are isomerized, built up from simple molecules, oxidized, hydrogenated, or caused to polymerize (polymerization). Only a few of the most important such processes and their catalysts can be mentioned here. Catalysts are of two physical types: homogeneous (i.e., dissolved in the reaction mixture) and heterogeneous (i.e., constituting a solid phase separate from and insoluble in the reaction mixture). Both types are represented on the industrial scene, but the latter are much more common.

      The introduction of catalysts that allow polymerization to be carried out at relatively low temperatures and pressures revolutionized the production of polyethylene and polypropylene. Previously polyethylene had to be made by a process requiring pressures of about 1,000 atmospheres, and polypropylene of useful properties was not commercially important. The catalysts devised and applied during the early 1950s are prepared from titanium tetrachloride and an aluminum alkyl such as triethylaluminum: the precipitated, titanium-containing solid, plus the excess aluminum alkyl in solution, constitutes the catalyst. A very different sort of catalyst, consisting of chromium trioxide dispersed on silica-alumina, performs similarly in polymerizing ethylene but cannot produce a useful form of polypropylene.

      Chromium in the form of chromium sesquioxide, or chromic oxide, on alumina is the major industrial catalyst for transforming saturated hydrocarbons (i.e., those in which all available valence bonds of the atoms are attached to other atoms) to useful olefins (unsaturated organic compounds), chiefly n-butane to butylene and butadiene.

      Iron-containing catalysts are used in various processes of which the most notable is that for producing ammonia from nitrogen and hydrogen. This process, developed early in the 20th century, represents the first major industrial application of transition metal catalysis. The catalyst is magnetic triiron tetroxide (Fe3O4), “promoted” by the addition of small quantities of potassium oxide, aluminum oxide, calcium oxide, and silica.

      Molybdenum in molybdenum trioxide–aluminum oxide mixtures or in cobalt oxide–molybdenum trioxide–aluminum oxide mixtures finds many applications, such as desulfurizing gases and liquids; removing lead, nickel, and vanadium from petroleum refinery feed stocks; and in modifying, or “reforming,” crude petroleum fractions to increase their octane rating.

      Olefins that are free of such impurities as carbon monoxide, sulfur, halogen, and compounds of arsenic or lead (catalyst poisons), can be hydrogenated (i.e., combined with hydrogen) at atmospheric pressure and room temperature, using various types of active nickel preparations as catalysts.

      Copper is a component of a variety of catalysts, of which the copper chromites, used for selective hydrogenation of carbonyl groups (compounds containing the organic radical C = O), are perhaps the most important. Fats and oils can also be hydrogenated to alcohols using copper catalysts. Palladium chloride together with salts of copper in the +1 oxidation state forms a homogeneous catalyst for oxidizing ethylene to acetaldehyde (Wacker process), but commercial use of this process has been abandoned. Metallic platinum has a broad spectrum of catalytic activities. One of the most important in terms of tonnage production is in catalytic reforming of petroleum fractions to improve antiknock quality of gasoline. Silver oxide, on an inactive, refractory support, catalyzes oxidation of ethylene to ethylene oxide.

      The oxo process (also called hydroformylation), in which certain olefins react with carbon monoxide and hydrogen to give aldehydes (aldehyde) (a class of organic compounds that contain the group −CHO), requires a homogeneous catalyst containing a transition metal. The first practicable one was hydrogen tetracarbonylcobaltate, HCo(CO)4, which is formed in the reaction mixture by action of hydrogen on octacarbonyldicobalt, Co2(CO)8. More recently rhodium complexes have been found to have greater activity at lower temperatures and pressures and to be more easily recovered. The net reaction in the oxo process is represented by

Biological functions of transition metals
      Several transition elements are important to the chemistry of living systems, the most familiar examples being iron, cobalt, copper, and molybdenum. Iron is by far the most widespread and important transition metal that has a function in living systems; proteins containing iron participate in two main processes, oxygen transport and electron transfer (i.e., oxidation–reduction) reactions. There are also a number of substances that act to store and transport iron itself.

      Though cobalt is understood to be an essential trace element in animal nutrition, the only detailed chemical knowledge of its biochemical action has to do with vitamin B12 and related co-enzymes. These molecules contain one atom of cobalt bound in a macrocyclic ring (i.e., one consisting of many atoms) called corrin, which is similar to a porphyrin ring. Copper is found in both plants and animals, and numerous copper-containing proteins have been isolated. The blood of many lower animals, such as mollusks, cephalopods, gastropods, and decapods, contains respiratory proteins called hemocyanins, which contain copper atoms (but no heme) and appear to bind one oxygen molecule per two copper atoms. Human serum contains a glycoprotein called ceruloplasmin, the molecule of which contains eight copper atoms; its biological function is still uncertain. Other proteins, called cerebrocuprein, erythrocuprein, and hepatocuprein, that are found in the mammalian brain, erythrocytes, and liver, respectively, contain about 60 percent of the total copper in those tissues; their functions are still unknown. There are a number of copper-containing enzymes; examples are (1) ascorbic acid oxidase (an oxidase is an oxidizing enzyme), which contains eight atoms of copper per molecule; it is widely distributed in plants and microorganisms; (2) cytochrome oxidase, which contains heme and copper in a 1:1 ratio; (3) tyrosinases, which catalyze the formation of melanin (brownish-black pigments occurring in hair, skin, and retina of higher animals) and were the first enzymes in which copper was shown to be essential to function.

      Vanadium occurs widely in petroleum, notably that from Venezuela, and can be isolated as porphyrin complexes, the origin of which is not known. Vanadium is present in high concentrations in blood cells (vanadocytes) of certain ascidians (sea squirts), apparently in a curious, complex, and poorly understood protein-containing substance called hemovanadin, thought to serve in oxygen transport. Molybdenum is believed to be a necessary trace element in animal diets, but its function and the minimum levels have not been established. Nitrogen-fixing bacteria utilize enzymes that contain both molybdenum and iron. One such enzyme, or at least a part of it that has been isolated in the crystalline state, contains two atoms of molybdenum and 40 atoms of iron. This protein in association with another, which contains only iron, can catalyze the reduction of nitrogen gas to nitrogen compounds.

      Efforts to understand the function of transition metals in biological systems have led to the growth of the field of bioinorganic chemistry.

Theories of transition-metal complexes
 As has been noted, partially filled orbitals account for the characteristic chemical properties of the regular transition elements, both as a class and as individuals. The interpretation and understanding of the chemical and physical properties of these elements thus depends heavily upon the description of these dn (n is one or more but fewer than ten) electron configurations. The five orbitals of each d shell, regardless of principal quantum number, have the shapes and designations shown in the Figure—>. The radial extent or size changes with principal quantum number, but the shapes are characteristic for all sets.

      For an atom or ion in free space and not subject to any electromagnetic field, the five orbitals of a set have equal energies; but in chemical compounds, the surrounding atoms cause the set to separate into subsets, differing in energy. This separation has a profound influence upon the spectroscopic properties (including colour) and magnetic properties of the substance, and the analysis of such effects has stimulated a great deal of theoretical and experimental effort. The most common situations to be considered are those in which a transition-metal ion occurs in an octahedron or a tetrahedron of surrounding anions or other ligands. Such cases illustrate the handling of the problem of the electronic structure of transition-metal ions in compounds.

      Three different theoretical approaches have been used: (1) the valence-bond treatment, pioneered in the United States by Linus Pauling; (2) the crystal-field theory or its more sophisticated form, the ligand-field theory (ligand field theory), first proposed by Hans Bethe and developed extensively by the U.S. physicist J.H. Van Vleck; and (3) the molecular orbital theory, the application of which to transition-metal complexes was first discussed by Van Vleck. The second and third methods are used almost exclusively, and only those two will be outlined here.

Crystal-field and ligand-field theories
      The crystal-field theory (CFT) employs an extreme electrostatic model, in which the ligands are treated as point charges (i.e., as if the entire negative charge were concentrated at a single point in space) if they are anions or as point dipoles (i.e., pairs of particles having equal and opposite charges that are separated by a finite distance) if they are neutral molecules. These extreme approximations are useful because they preserve certain essential features of the actual physical condition and yet reduce the mathematical problem to one that can be solved. Since the CFT treatment makes no allowance for covalence in the metal–ligand bonds, it necessarily does not account for all aspects of the electronic structures of complexes, even the most ionic ones. It can be amended, however, by empirical adjustment of certain parameters, to allow for some of the effects of covalence, without sacrificing mathematical convenience. This version of the theory, which can correlate much of the data on complexes of metals in their normal oxidation states, is called ligand-field theory (LFT).

Orbital splitting patterns
  Crystal-field theory treats the question of how the energies of a set of d electrons or orbitals are split when a set of ligands is placed around a central metal ion; it does so by treating the ligands as a set of negative charges. The simplest case to consider is that of an ion with a single d electron, surrounded by six negative charges at the vertices of an octahedron. This arrangement is defined, relative to a set of Cartesian axes, x, y, z, shown in the Figure—>. By comparing the shapes of the d orbitals (see Figure—>) with this arrangement, it can be seen that the dxy, dxz, and dyz orbitals have equivalent relationships to the set of charges. Thus, the electron will be repelled to the same extent by the negative charges regardless of which of the three orbitals it occupies. The three orbitals thus have equal energy and are called triply degenerate. It is not particularly obvious from a pictorial argument, but mathematical analysis shows that each of the other two orbitals, dz2 and dx2− y2, causes the electron to experience the same amount of electrostatic repulsion from the surrounding charges; they are described as doubly degenerate. Thus, the first important conclusion of the crystal-field theory is that the spatial relationships of the d orbitals to the surrounding charges cause the set of five d orbitals to be split into two subsets; the orbitals within each subset are equivalent to each other and are thus degenerate, but they are no longer degenerate with those in the other subset. Those in the subset consisting of the dxy, dyz, and dzx orbitals are known as the t2g orbitals; the dz2 and dx2− y2 pair are called the eg orbitals. When the surrounding charges are located at the vertices of an octahedron, the two eg orbitals are of higher energy than the three t2g orbitals. When an ion is surrounded by a tetrahedrally arranged set of four negative charges, the d orbitals also split into a set of three and a set of two, but the energy order is the reverse of that in the octahedral case. Beginning with these splitting patterns it is possible to elaborate a detailed account of the magnetic and spectroscopic properties of transition-metal ions in their compounds.

Magnetic (magnetism) and spectroscopic properties
      Discussion of magnetic properties must begin with the basic question of how many unpaired electrons will be present. This is decided by the competition between two factors: (1) the electrons tend to occupy separate orbitals since this minimizes repulsions between them; and (2) the electrons also tend to occupy orbitals of lower energy in preference to higher ones. For ions with one, two, and three d electrons in octahedral fields, the first factor dictates, without opposition from the second, that the electrons occupy separate t2g orbitals; such ions therefore always have 1, 2, and 3 unpaired electrons. Similarly, for configurations of eight and nine d electrons, the only possibilities are six electrons in t2g orbitals and two in eg orbitals, and six electrons in t2g orbitals and three in eg orbitals (t2g6eg2 and t2g6eg3), and such ions must always have 2 and 1 unpaired electrons, respectively. For cases of d4, d5, d6, and d7 configurations, the number of unpaired electrons depends on which of the two factors dominates. If the splitting (the energy difference between eg and t2g orbitals) is relatively small, control by the first factor produces the high-spin configurations, t2g3eg, t2g3eg2, t2g4eg2, and t2g5eg2, which have 4, 5, 4, and 3 unpaired electrons, respectively. When the difference is relatively large, control by the second factor produces the low-spin configurations t2g4, t2g5, t2g6, and t2g6eg, with 2, 1, 0, and 1 unpaired electrons, respectively. In tetrahedral complexes the splitting is always sufficiently small so that the first factor dominates and high-spin states are always obtained. Once the number of unpaired electrons is known, numerical values of the magnetic moments are calculated by taking account of the interaction of the orbital-angular momentum with the spin-angular momentum.

      The spectra associated with the electronic structures of transition-metal ions, which are responsible for their colours, can be understood in terms of the CFT splitting patterns. For a d1 ion in an octahedral field, a single electronic transition, t2geg, is expected; that is, the absorption of light raises the energy of an electron and causes it to pass from the low-energy t2g orbital to the high-energy eg orbital. The titanium ion Ti3+, for example, has just a single absorption band in the visible region of the spectrum, at a wavelength of about 5,000 nanometres, corresponding to an energy of 20,000 cm−1, which is assigned to this transition. This says that the orbital energy difference for Ti3+ in [Ti(H2O)6]3+ is 20,000 per centimetre. For configurations with more than one d electron, electronic interactions require that an elaborate treatment, which cannot be explained here, be used.

      The difference in the colours of hexaquonickel ion, [Ni(H2O)6]2+ (green), and tris (ethylenediamine) nickel ion, [Ni(H2NCH2CH2NH2) 3]2+ (purple), reflects the fact that the six nitrogen atoms cause a greater splitting than the six oxygen atoms.

      In general, the relative magnitudes of d orbital splittings for a given ion with different ligand sets fall in a consistent order. This ordering of ligands according to their ability to split the d orbitals was discovered empirically in the 1930s and is called the spectrochemical series. A short list of ligands in order of their splitting power is fluoride (weakest), water, thiocyanate, pyridine or ammonia, ethylenediamine, nitrite, cyanide (strongest).

Jahn-Teller effect
      According to the Jahn-Teller theorem, any molecule or complex ion in an electronically degenerate state will be unstable relative to a configuration of lower symmetry in which the degeneracy is absent. The chief applications of this theorem in transition-metal chemistry are in connection with octahedrally coordinated metal ions with high-spin d4, low-spin d7, and d9 configurations; in each of these cases the t2g orbitals are all equally occupied (either all half filled or all filled) and there is a single electron or a single vacancy in the eg orbitals. Either an eg or an eg3 configuration gives rise to a doubly degenerate (E) ground state, and thus a distortion of the octahedron is expected. In other words, high-spin d4, low-spin d7, and d9 ions should be found in distorted, not regular, octahedral environments. It has been found by experiment that, with few possible exceptions, this is the case. The exceptions do not necessarily constitute violations of the Jahn-Teller theorem, because the slightness of the distortions may be within experimental error.

      By far the most common form of distortion in the three cases given above is an elongation of the octahedron along one of its fourfold axes. Thus, two opposite ligands are farther from the metal ion than the other, coplanar (i.e., lying in one plane) set of four. Such effects have been observed in compounds and complexes of the chromium ion Cr2+ and the manganese ion Mn3+ in high-spin configurations (t2g3eg); the cobalt ion Co2+ and the nickel ion Ni3+ in low-spin configurations (t2g6eg); and the copper ion Cu2+ and the silver ion Ag+ (t2g6eg3). The greatest body of data available is for compounds of copper in the +2 oxidation state, and some representative sets of copper–ligand distances are copper chloride, four chloride ions at 2.30 Å and two at 2.95 Å; copper bromide, four bromide ions at 2.40 Å and two at 3.18 Å.

Molecular-orbital theory
      The molecular-orbital (MO) treatment of the electronic structures of transition-metal complexes is, in principle, a more flexible approach than the CFT or LFT treatments. Because a great many complexes and compounds in the ordinary oxidation states (+2, +3) of the transition metals are substantially ionic, the CFT and LFT treatments are useful though not exact. In the MO method, the ligand-atom orbitals and all of the valence-shell orbitals of the metal atom are presumed to interact to form molecular orbitals.

      The interaction of a metal ion from the first transition series and an octahedral set of ligands that interact only through orbitals directed toward the metal ion results in the separation of the d orbitals of the metal ion into two sets of molecular orbitals; a t2g set and a higher energy eg set. This is the same pattern as that obtained in the CFT treatment, but the difference is that now the t2g and eg orbitals are regarded not as pure d orbitals but as molecular orbitals, having both metal d orbital and ligand orbital components. Thus, the modifications of CFT that lead to LFT—that is, inclusion of covalence effects—are in harmony with the basic ideas of the MO treatment.

The elements of the first transition series
      Although the transition elements have many general chemical similarities, each one has a detailed chemistry of its own. The closest relationships are usually to be found among the three elements in each vertical group in the periodic table, although within each group the element of the first series usually differs more from the other two than they differ from each other. Most of the first series elements are more familiar and technically important than the heavier members of their vertical group.

      A few of the chemical trends to be found in the first transition series may be capsulized.

1. From titanium to manganese the highest oxidation state exhibited, which usually is found only in oxo compounds, fluorides, or chlorides, corresponds to the total number of 3d and 4s electrons in the atom. The stability of this highest oxidation state (oxidation number) decreases from titanium in the +4 state to manganese in the +7 state. Following manganese—that is, for iron, cobalt, and nickel—oxidation states corresponding to the loss of all 3d and 4s electrons do not occur; higher oxidation states in general become progressively more difficult to attain because the increasing nuclear charge causes the 3d electrons to be more tightly bound. Very high oxidation states occur only for chromium (+5, +6 states), manganese (+5, +6, +7 states), and iron (+5, +6 states) and apart from the fluorides, such as chromium pentafluoride, CrF5 (with chromium in the +5 state), and chromium hexafluoride, CrF6 (with chromium in the +6 state), and oxofluorides such as manganese trioxide fluoride, MnO3F (with manganese in the +7 state), the main chemistry in these oxidation states is that of oxo anions such as permanganate, MnO4 (+7 state); chromate, CrO42− (+6 state); and ferrate, FeO42− (+6 state). All of these compounds are powerful oxidizing agents.
2. The oxides (oxide) of each element become more acidic with increasing oxidation number, and the halides become more covalent and susceptible to hydrolysis.
3. In the oxo anions characteristic of the higher oxidation states the metal atom is tetrahedrally surrounded by oxygen atoms, whereas in the oxides formed in the lower oxidation states the atoms are usually octahedrally coordinated.
4. In the oxidation states +2 and +3, complexes in aqueous solution or in crystals are usually four-, five- or six-coordinated.
5. Oxidation states lower than +2 are not found in the ordinary chemistries of the transition elements, except for copper. The lower oxidation states are, however, attainable for all the elements using ligands of the carbon monoxide type.

The elements of the second and third transition series
      While the elements in the second and third transition series for a given group have chemical properties similar to those of the element in the first series, they nevertheless show definite differences from the lighter element of the group. The following examples illustrate this point: (1) Although cobalt (of the first series) forms a considerable number of tetrahedral and octahedral complexes in its +2 oxidation state, and that state is characteristic in ordinary aqueous chemistry, the +2 states of rhodium (second series) and iridium (third series) are rare and relatively unimportant. (2) The manganese ion Mn2+ is very stable and of principal importance in the chemistry of manganese, but for technetium and rhenium the oxidation state +2 is little more than a laboratory curiosity. (3) Chromium in its +3 state forms a great number of complexes, which make up one of the best known aspects of the chemistry of the element; whereas the +3 states of molybdenum and tungsten are not particularly stable states under any conditions and form only a few complexes. (4) The oxo anions of the first-row elements in their higher oxidation states—for example, chromate (with chromium in the +6 state) and permanganate (manganese, +7 state)—are powerful oxidizing agents the chemistry of which is essentially restricted to that function; whereas their stoichiometric analogues, such as molybdate (molybdenum, +6 state), tungstate (tungsten, +6 state), pertechnetate (technetium, +7 state), and perrhenate (rhenium, +7 state), are quite stable and have an extensive and diverse chemistry. There are, however, some cases in which quite valid and useful analogies can be found between the chemistry of the lighter element and the two heavier elements of the group. For instance, the chemistry of complexes of rhodium in the +3 state is, in general, quite similar to that of complexes of cobalt in the +3 state. On the whole, however, there are differences more consistently than there are similarities between the first element and the heavier elements of each group.

      For the heavier transition elements, higher oxidation states are generally more stable than is the case for the elements in the first transition series; this is true not only, as has been mentioned, for the properties of the oxo anions but for the higher halides as well. Thus the heavier elements form compounds such as ruthenium oxide, RuO4 (+8 state); tungsten chloride, WCl6 (+6 state); platinum fluoride, PtF6 (+6 state), etc., which have no analogues among the first-row elements, whereas the chemistry of aquo ions of lower oxidation states, especially +2 and +3, which is such a dominant part of the chemistry of the lighter elements, is relatively unimportant for most of the heavier ones.

      One of the reasons for the generally close similarity in chemistry between the second and third transition series elements of a given group is the so-called lanthanide contraction. As already described, the series of elements known as the lanthanoids comes between the second and third regular transition series and is formed by the filling of the 4f orbitals. The 5d and 6s orbitals, which are the valence-shell orbitals for the third transition series elements, are imperfectly shielded by the 4f electrons from the increased nuclear charge. There is consequently, a steady contraction in the size of these orbitals through the lanthanoid series of elements, with the net result that the atomic and ionic radii of hafnium, which immediately follows the lanthanoid series, are almost identical to the corresponding radii of the zirconium atom, which lies just above it in Group 4 (IVb). Because the zirconium and hafnium atoms have almost identical sizes as well as analogous electron configurations in all of their oxidation states, their chemistry and the properties of their compounds are exceedingly similar; indeed, it is very difficult to separate the two elements because of the great similarity in properties of their compounds. In moving along the third transition series, there is a steady but slow divergence in the properties of the second and third elements in each group so that, at the end, considerable differences exist between those of palladium and platinum and of silver and gold. The differences are not as great, however, as might have been expected had the lanthanide contraction not intervened to preclude greater disparity in the orbital sizes. Although niobium and tantalum are not quite as similar as zirconium and hafnium, the differences between them are slight, and, similarly, molybdenum and tungsten, technetium and rhenium, ruthenium and osmium, and rhodium and iridium show marked similarities in their chemistries.

F. Albert Cotton

Additional Reading
Edwin M. Larsen, Transitional Elements (1965), gives a brief and elementary topical discussion. D.L. Kepert, The Early Transition Metals (1972), presents a detailed discussion of the chemistry of the metals in groups 4, 5, and 6. J.P. Candlin, K.A. Taylor, and D.T. Thompson, Reactions of Transition-Metal Complexes (1968), covers its subject matter in great detail. Also useful is Geoffrey Wilkinson, Robert D. Gillard, and Jon A. McCleverty (eds.), Comprehensive Coordination Chemistry (1987). Charles L. Thomas, Catalytic Processes and Proven Catalysts (1970), reviews the actual uses of industrial catalysts by process type. Carl J. Ballhausen, Introduction to Ligand Field Theory (1962), gives a working account of how ligand field theory is applied to the d-block elements. The electronic structures of transition-metal compounds and complexes are covered in Carl J. Ballhausen and Harry B. Gray, Molecular Electronic Structures: An Introduction (1980); Carl J. Ballhausen, Molecular Electronic Structures of Transition Metal Complexes (1979); and D.T. Burns, A. Townshend, and A.G. Catchpole, Inorganic Reaction Chemistry, vol. 2, Reactions of the Elements and Their Compounds, 2 parts (1981). Also of interest are W.P. Griffith, The Chemistry of the Rarer Platinum Metals (1967); R.J.H. Clark, The Chemistry of Titanium and Vanadium (1968); R. Colton, The Chemistry of Rhenium and Technetium (1965); F. Fairbrother, The Chemistry of Niobium and Tantalum (1967); C.T. Horovitz (ed.), Scandium: Its Occurrence, Chemistry, Physics, Metallurgy, Biology, and Technology (1975); Elaine A. Seddon and Kenneth R. Seddon, The Chemistry of Ruthenium (1984); Richard J. Puddephatt, The Chemistry of Gold (1978); and S.A. Cotton and Frank A. Hart, The Heavy Transition Elements (1975).F. Albert Cotton Ed.

* * *


Universalium. 2010.

Игры ⚽ Поможем решить контрольную работу

Look at other dictionaries:

  • transition element — n. any element of several groups of elements formed by adding electrons to an inner shell as the atomic number increases …   English World dictionary

  • transition element — noun or transition metal : any of the series of metals (as scandium, titanium, vanadium, chromium, manganese, iron, cobalt, nickel) that fall in the center of the long form of the periodic table, that include as inner series the lanthanide series …   Useful english dictionary

  • transition element — pereinamasis elementas statusas T sritis fizika atitikmenys: angl. transition element; transitional element vok. Übergangselement, n rus. переходный элемент, m pranc. élément de transition, m …   Fizikos terminų žodynas

  • transition element — noun see transition metal …   New Collegiate Dictionary

  • transition element — noun any of the group or metallic elements that have an incomplete inner (d) electron shell; they have multiple valencies and form coloured compounds and stable complex ions; the elements …   Wiktionary

  • transition element — /trænˈzɪʃən ɛləmənt/ (say tran zishuhn eluhmuhnt) noun any of numerous metallic elements in which the two outermost electron shells are incomplete; they have variable valencies and coloured ions. Also, transition metal …  

  • transition element — transi′tion el ement n. chem. any element in the four series of chemical elements with atomic numbers 21–29, 39–47, 57–79, and 89–107, that in a given inner orbital has less than a full quota of electrons. Also called transi′tion met al •… …   From formal English to slang

  • переходный элемент ( transition element) — 3.4 переходный элемент ( transition element): Элемент воздуховода, подогнанный к воздуховоду источника звука и соединяющий его с измерительным воздуховодом и в некоторых случаях соединяющий измерительный воздуховод с глушителем. Источник …   Словарь-справочник терминов нормативно-технической документации

  • transition — [ trɑ̃zisjɔ̃ ] n. f. • 1501 « procédé rhétorique »; lat. transitio, proprt « passage » 1 ♦ Manière de passer de l expression d une idée à une autre, de lier les parties d un discours. ⇒ passage. L art des transitions. Ménager les transitions.… …   Encyclopédie Universelle

  • Transition metal — The term transition metal (sometimes also called a transition element) has two possible meanings: The IUPAC definition[1] states that a transition metal is an element whose atom has an incomplete d sub shell, or which can give rise to cations… …   Wikipedia

Share the article and excerpts

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