Spin states (d electrons)

Last updated

Spin states when describing transition metal coordination complexes refers to the potential spin configurations of the central metal's d electrons. For several oxidation states, metals can adopt high-spin and low-spin configurations. The ambiguity only applies to first row metals, because second- and third-row metals are invariably low-spin. These configurations can be understood through the two major models used to describe coordination complexes; crystal field theory and ligand field theory (a more advanced version based on molecular orbital theory). [1]

Contents

High-spin vs. low-spin

Octahedral complexes

Low-spin [Fe(NO2)6] crystal field diagram CFT-Low Spin Splitting Diagram-Vector.svg
Low-spin [Fe(NO2)6] crystal field diagram

The Δ splitting of the d orbitals plays an important role in the electron spin state of a coordination complex. Three factors affect Δ: the period (row in periodic table) of the metal ion, the charge of the metal ion, and the field strength of the complex's ligands as described by the spectrochemical series. Only octahedral complexes of first row transition metals adopt high-spin states.

In order for low spin splitting to occur, the energy cost of placing an electron into an already singly occupied orbital must be less than the cost of placing the additional electron into an eg orbital at an energy cost of Δ. If the energy required to pair two electrons is greater than the energy cost of placing an electron in an eg, Δ, high spin splitting occurs.

If the separation between the orbitals is large, then the lower energy orbitals are completely filled before population of the higher orbitals according to the Aufbau principle. Complexes such as this are called "low-spin" since filling an orbital matches electrons and reduces the total electron spin. If the separation between the orbitals is small enough then it is easier to put electrons into the higher energy orbitals than it is to put two into the same low-energy orbital, because of the repulsion resulting from matching two electrons in the same orbital. So, one electron is put into each of the five d orbitals before any pairing occurs in accord with Hund's rule resulting in what is known as a "high-spin" complex. Complexes such as this are called "high-spin" since populating the upper orbital avoids matches between electrons with opposite spin.

High-spin [FeBr6] crystal field diagram CFT-High Spin Splitting Diagram-Vector.svg
High-spin [FeBr6] crystal field diagram

The charge of the metal center plays a role in the ligand field and the Δ splitting. The higher the oxidation state of the metal, the stronger the ligand field that is created. In the event that there are two metals with the same d electron configuration, the one with the higher oxidation state is more likely to be low spin than the one with the lower oxidation state; for example, Fe2+ and Co3+ are both d6; however, the higher charge of Co3+ creates a stronger ligand field than Fe2+. All other things being equal, Fe2+ is more likely to be high spin than Co3+.

Ligands also affect the magnitude of Δ splitting of the d orbitals according to their field strength as described by the spectrochemical series. Strong-field ligands, such as CN and CO, increase the Δ splitting and are more likely to be low-spin. Weak-field ligands, such as I and Br cause a smaller Δ splitting and are more likely to be high-spin.

Some octahedral complexes exhibit spin crossover, where the high and low spin states exist is dynamic equilibrium.

Light-induced spin-crossover of [Fe(pyCH2NH2)3] , which switches from high and low-spin. Fe(picNH2)3.png
Light-induced spin-crossover of [Fe(pyCH2NH2)3] , which switches from high and low-spin.

Tetrahedral complexes

Fe(4-norbornyl)4 is a rare example of a low-spin tetrahedral complex. Fe(4-norbornyl)4.svg
Fe(4-norbornyl)4 is a rare example of a low-spin tetrahedral complex.

The Δ splitting energy for tetrahedral metal complexes (four ligands), Δtet is smaller than that for an octahedral complex. Consequently, tetrahedral complexes are almost always high spin [3] Examples of low spin tetrahedral complexes include Fe(2-norbornyl)4, [4] [Co(4-norbornyl)4]+, and the nitrosyl complex Cr(NO)((N(tms)2)3.

Square planar complexes

Many d8 complexes of the first row metals exist in tetrahedral or square planar geometry. In some cases these geometries exist in measurable equilibria. For example, dichlorobis(triphenylphosphine)nickel(II) has been crystallized in both tetrahedral and square planar geometries. [5]

Ligand field theory vs crystal field theory

In terms of d-orbital splitting, ligand field theory (LFT) and crystal field theory (CFT) give similar results. CFT is an older, simpler model that treats ligands as point charges. LFT is more chemical, emphasizes covalent bonding and accommodates pi-bonding explicitly.

High-spin and low-spin systems

In the case of octahedral complexes, the question of high spin vs low spin first arises for d4, since it has more than the 3 electrons to fill the non-bonding d orbitals according to ligand field theory or the stabilized d orbitals according to crystal field splitting.

All complexes of second and third row metals are low-spin.

d4
Octahedral high-spin: 4 unpaired electrons, paramagnetic, substitutionally labile. Includes Cr2+ (many complexes assigned as Cr(II) are however Cr(III) with reduced ligands [6] ), Mn3+.
Octahedral low-spin: 2 unpaired electrons, paramagnetic, substitutionally inert. Includes Cr2+, Mn3+.
d5
Octahedral high-spin: 5 unpaired electrons, paramagnetic, substitutionally labile. Includes Fe3+, Mn2+. Example: Tris(acetylacetonato)iron(III).
Octahedral low-spin: 1 unpaired electron, paramagnetic, substitutionally inert. Includes Fe3+. Example: [Fe(CN)6]3−.
d6
Octahedral high-spin: 4 unpaired electrons, paramagnetic, substitutionally labile. Includes Fe2+, Co3+. Examples: [Fe(H2O)6]2+, [CoF6]3−.
Octahedral low-spin: no unpaired electrons, diamagnetic, substitutionally inert. Includes Fe2+, Co3+, Ni4+. Example: [Co(NH3)6]3+.
d7
Octahedral high-spin: 3 unpaired electrons, paramagnetic, substitutionally labile. Includes Co2+, Ni3+.
Octahedral low-spin:1 unpaired electron, paramagnetic, substitutionally labile. Includes Co2+, Ni3+. Example: [Co(NH3)6]2+.
d8
Octahedral high-spin: 2 unpaired electrons, paramagnetic, substitutionally labile. Includes Ni2+. Example: [Ni(NH3)6]2+.
Tetrahedral high-spin: 2 unpaired electrons, paramagnetic, substitutionally labile. Includes Ni2+. Example: [NiCl4]2-.
Square planar low-spin: no unpaired electrons, diamagnetic, substitutionally inert. Includes Ni2+. Example: [Ni(CN)4]2−.

Ionic radii

The spin state of the complex affects an atom's ionic radius. For a given d-electron count, high-spin complexes are larger. [7]

d4
Octahedral high spin: Cr2+, 64.5 pm.
Octahedral low spin: Mn3+, 58 pm.
d5
Octahedral high spin: Fe3+, the ionic radius is 64.5 pm.
Octahedral low spin: Fe3+, the ionic radius is 55 pm.
d6
Octahedral high spin: Fe2+, the ionic radius is 78 pm, Co3+ ionic radius 61 pm.
Octahedral low spin: Includes Fe2+ ionic radius 62 pm, Co3+ ionic radius 54.5 pm, Ni4+ ionic radius 48 pm.
d7
Octahedral high spin: Co2+ ionic radius 74.5 pm, Ni3+ ionic radius 60 pm.
Octahedral low spin: Co2+ ionic radius 65 pm, Ni3+ionic radius 56 pm.
d8
Octahedral high spin: Ni2+ ionic radius 69 pm.
Square planar low-spin: Ni2+ ionic radius 49 pm.

Ligand exchange rates

Generally, the rates of ligand dissociation from low spin complexes are lower than dissociation rates from high spin complexes. In the case of octahedral complexes, electrons in the eg levels are anti-bonding with respect to the metal-ligand bonds. Famous "exchange inert" complexes are octahedral complexes of d3 and low-spin d6 metal ions, illustrated respectfully by Cr3+ and Co3+. [8]

Related Research Articles

<span class="mw-page-title-main">Coordination complex</span> Molecule or ion containing ligands datively bonded to a central metallic atom

A coordination complex is a chemical compound consisting of a central atom or ion, which is usually metallic and is called the coordination centre, and a surrounding array of bound molecules or ions, that are in turn known as ligands or complexing agents. Many metal-containing compounds, especially those that include transition metals, are coordination complexes.

<span class="mw-page-title-main">Ligand</span> Ion or molecule that binds to a central metal atom to form a coordination complex

In coordination chemistry, a ligand is an ion or molecule with a functional group that binds to a central metal atom to form a coordination complex. The bonding with the metal generally involves formal donation of one or more of the ligand's electron pairs, often through Lewis bases. The nature of metal–ligand bonding can range from covalent to ionic. Furthermore, the metal–ligand bond order can range from one to three. Ligands are viewed as Lewis bases, although rare cases are known to involve Lewis acidic "ligands".

<span class="mw-page-title-main">Paramagnetism</span> Weak, attractive magnetism possessed by most elements and some compounds

Paramagnetism is a form of magnetism whereby some materials are weakly attracted by an externally applied magnetic field, and form internal, induced magnetic fields in the direction of the applied magnetic field. In contrast with this behavior, diamagnetic materials are repelled by magnetic fields and form induced magnetic fields in the direction opposite to that of the applied magnetic field. Paramagnetic materials include most chemical elements and some compounds; they have a relative magnetic permeability slightly greater than 1 and hence are attracted to magnetic fields. The magnetic moment induced by the applied field is linear in the field strength and rather weak. It typically requires a sensitive analytical balance to detect the effect and modern measurements on paramagnetic materials are often conducted with a SQUID magnetometer.

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

<span class="mw-page-title-main">Ferric</span> The element iron in its +3 oxidation state

In chemistry, Ferric refers to the element iron in its +3 oxidation state. Ferric chloride is an alternative name for iron(III) chloride (FeCl3). The adjective ferrous is used instead for iron(II) salts, containing the cation Fe2+. The word ferric is derived from the Latin word ferrum, meaning "iron".

<span class="mw-page-title-main">Ferrous</span> The element iron in its +2 oxidation state

In chemistry, iron(II) refers to the element iron in its +2 oxidation state. The adjective ferrous or the prefix ferro- is often used to specify such compounds, as in ferrous chloride for iron(II) chloride (FeCl2). The adjective ferric is used instead for iron(III) salts, containing the cation Fe3+. The word ferrous is derived from the Latin word ferrum, meaning "iron".

In molecular physics, crystal field theory (CFT) describes the breaking of degeneracies of electron orbital states, usually d or f orbitals, due to a static electric field produced by a surrounding charge distribution. This theory has been used to describe various spectroscopies of transition metal coordination complexes, in particular optical spectra (colors). CFT successfully accounts for some magnetic properties, colors, hydration enthalpies, and spinel structures of transition metal complexes, but it does not attempt to describe bonding. CFT was developed by physicists Hans Bethe and John Hasbrouck van Vleck in the 1930s. CFT was subsequently combined with molecular orbital theory to form the more realistic and complex ligand field theory (LFT), which delivers insight into the process of chemical bonding in transition metal complexes. CFT can be complicated further by breaking assumptions made of relative metal and ligand orbital energies, requiring the use of inverted ligand field theory (ILFT) to better describe bonding.

Ligand field theory (LFT) describes the bonding, orbital arrangement, and other characteristics of coordination complexes. It represents an application of molecular orbital theory to transition metal complexes. A transition metal ion has nine valence atomic orbitals - consisting of five nd, one (n+1)s, and three (n+1)p orbitals. These orbitals have the appropriate energy to form bonding interactions with ligands. The LFT analysis is highly dependent on the geometry of the complex, but most explanations begin by describing octahedral complexes, where six ligands coordinate with the metal. Other complexes can be described with reference to crystal field theory. Inverted ligand field theory (ILFT) elaborates on LFT by breaking assumptions made about relative metal and ligand orbital energies.

The Jahn–Teller effect is an important mechanism of spontaneous symmetry breaking in molecular and solid-state systems which has far-reaching consequences in different fields, and is responsible for a variety of phenomena in spectroscopy, stereochemistry, crystal chemistry, molecular and solid-state physics, and materials science. The effect is named for Hermann Arthur Jahn and Edward Teller, who first reported studies about it in 1937.

<span class="mw-page-title-main">Octahedral molecular geometry</span> Molecular geometry

In chemistry, octahedral molecular geometry, also called square bipyramidal, describes the shape of compounds with six atoms or groups of atoms or ligands symmetrically arranged around a central atom, defining the vertices of an octahedron. The octahedron has eight faces, hence the prefix octa. The octahedron is one of the Platonic solids, although octahedral molecules typically have an atom in their centre and no bonds between the ligand atoms. A perfect octahedron belongs to the point group Oh. Examples of octahedral compounds are sulfur hexafluoride SF6 and molybdenum hexacarbonyl Mo(CO)6. The term "octahedral" is used somewhat loosely by chemists, focusing on the geometry of the bonds to the central atom and not considering differences among the ligands themselves. For example, [Co(NH3)6]3+, which is not octahedral in the mathematical sense due to the orientation of the N−H bonds, is referred to as octahedral.

A spectrochemical series is a list of ligands ordered by ligand "strength", and a list of metal ions based on oxidation number, group and element. For a metal ion, the ligands modify the difference in energy Δ between the d orbitals, called the ligand-field splitting parameter in ligand field theory, or the crystal-field splitting parameter in crystal field theory. The splitting parameter is reflected in the ion's electronic and magnetic properties such as its spin state, and optical properties such as its color and absorption spectrum.

Copper proteins are proteins that contain one or more copper ions as prosthetic groups. Copper proteins are found in all forms of air-breathing life. These proteins are usually associated with electron-transfer with or without the involvement of oxygen (O2). Some organisms even use copper proteins to carry oxygen instead of iron proteins. A prominent copper protein in humans is in cytochrome c oxidase (cco). This enzyme cco mediates the controlled combustion that produces ATP. Other copper proteins include some superoxide dismutases used in defense against free radicals, peptidyl-α-monooxygenase for the production of hormones, and tyrosinase, which affects skin pigmentation.

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

Siegenite (also called grimmite, or nickel cobalt sulfide) is a ternary transition metal dichalcogenide compound with the chemical formula (Ni,Co)3S4. It has been actively studied as a promising material system for electrodes in electrochemical energy applications due to its better conductivity, greater mechanical and thermal stability, and higher performance compared to metal oxides currently in use. Potential applications of this material system include supercapacitors, batteries, electrocatalysis, dye-sensitized solar cells, photocatalysis, glucose sensors, and microwave absorption.

The 18-electron rule is a chemical rule of thumb used primarily for predicting and rationalizing formulas for stable transition metal complexes, especially organometallic compounds. The rule is based on the fact that the valence orbitals in the electron configuration of transition metals consist of five (n−1)d orbitals, one ns orbital, and three np orbitals, where n is the principal quantum number. These orbitals can collectively accommodate 18 electrons as either bonding or non-bonding electron pairs. This means that the combination of these nine atomic orbitals with ligand orbitals creates nine molecular orbitals that are either metal-ligand bonding or non-bonding. When a metal complex has 18 valence electrons, it is said to have achieved the same electron configuration as the noble gas in the period, lending stability to the complex. Transition metal complexes that deviate from the rule are often interesting or useful because they tend to be more reactive. The rule is not helpful for complexes of metals that are not transition metals. The rule was first proposed by American chemist Irving Langmuir in 1921.

In coordination chemistry, Tanabe–Sugano diagrams are used to predict absorptions in the ultraviolet (UV), visible and infrared (IR) electromagnetic spectrum of coordination compounds. The results from a Tanabe–Sugano diagram analysis of a metal complex can also be compared to experimental spectroscopic data. They are qualitatively useful and can be used to approximate the value of 10Dq, the ligand field splitting energy. Tanabe–Sugano diagrams can be used for both high spin and low spin complexes, unlike Orgel diagrams, which apply only to high spin complexes. Tanabe–Sugano diagrams can also be used to predict the size of the ligand field necessary to cause high-spin to low-spin transitions.

An electric effect influences the structure, reactivity, or properties of a molecule but is neither a traditional bond nor a steric effect. In organic chemistry, the term stereoelectronic effect is also used to emphasize the relation between the electronic structure and the geometry (stereochemistry) of a molecule.

The spin transition is an example of transition between two electronic states in molecular chemistry. The ability of an electron to transit from a stable to another stable electronic state in a reversible and detectable fashion, makes these molecular systems appealing in the field of molecular electronics.

The d electron count or number of d electrons is a chemistry formalism used to describe the electron configuration of the valence electrons of a transition metal center in a coordination complex. The d electron count is an effective way to understand the geometry and reactivity of transition metal complexes. The formalism has been incorporated into the two major models used to describe coordination complexes; crystal field theory and ligand field theory, which is a more advanced version based on molecular orbital theory. However the d electron count of an atom in a complex is often different from the d electron count of a free atom or a free ion of the same element.

Magnetochemistry is concerned with the magnetic properties of chemical compounds. Magnetic properties arise from the spin and orbital angular momentum of the electrons contained in a compound. Compounds are diamagnetic when they contain no unpaired electrons. Molecular compounds that contain one or more unpaired electrons are paramagnetic. The magnitude of the paramagnetism is expressed as an effective magnetic moment, μeff. For first-row transition metals the magnitude of μeff is, to a first approximation, a simple function of the number of unpaired electrons, the spin-only formula. In general, spin–orbit coupling causes μeff to deviate from the spin-only formula. For the heavier transition metals, lanthanides and actinides, spin–orbit coupling cannot be ignored. Exchange interaction can occur in clusters and infinite lattices, resulting in ferromagnetism, antiferromagnetism or ferrimagnetism depending on the relative orientations of the individual spins.

The spinels are any of a class of minerals of general formulation AB
2
X
4
which crystallise in the cubic (isometric) crystal system, with the X anions arranged in a cubic close-packed lattice and the cations A and B occupying some or all of the octahedral and tetrahedral sites in the lattice. Although the charges of A and B in the prototypical spinel structure are +2 and +3, respectively, other combinations incorporating divalent, trivalent, or tetravalent cations, including magnesium, zinc, iron, manganese, aluminium, chromium, titanium, and silicon, are also possible. The anion is normally oxygen; when other chalcogenides constitute the anion sublattice the structure is referred to as a thiospinel.

References

  1. Miessler, Gary L.; Donald A. Tarr (1998). Inorganic Chemistry (2nd ed.). Upper Saddle River, New Jersey: Pearson Education, Inc. Pearson Prentice Hall. ISBN   0-13-841891-8.
  2. Gütlich, P. (2001). "Photoswitchable Coordination Compounds". Coordination Chemistry Reviews. 219–221: 839–879. doi:10.1016/S0010-8545(01)00381-2.
  3. Zumdahl, Steven (2009). "19.6 Transition Metals and Coordination Chemistry: The Crystal Field Model". Chemical Principles. Cengage Learning, Inc. ISBN   978-0538734561.
  4. Bower, Barton K.; Tennent, Howard G. (1972). "Transition Metal Bicyclo[2.2.1]hept-1-yls". Journal of the American Chemical Society. 94 (7): 2512–2514. doi:10.1021/ja00762a056.
  5. Batsanov, Andrei S.; Howard, Judith A. K. (2001). "trans-Dichlorobis(triphenylphosphine)nickel(II) Bis(dichloromethane) Solvate: Redetermination at 120 K". Acta Crystallogr E. 57: 308–309. doi:10.1107/S1600536801008741. S2CID   97381117.
  6. Scarborough, Christopher C.; Sproules, Stephen; Doonan, Christian J.; Hagen, Karl S.; Weyhermüller, Thomas; Wieghardt, Karl (2012). "Scrutinizing Low-Spin Cr(II) Complexes". Inorganic Chemistry. 51 (12): 6969–6982. doi:10.1021/ic300882r. PMID   22676275.
  7. Shannon R.D. (1976). "Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides". Acta Crystallographica. A32 (5): 751–767. doi:10.1107/S0567739476001551.
  8. R. G. Wilkins (1991). Kinetics and Mechanism of Reactions of Transition Metal Complexes, 2nd Thoroughly Revised Edition. Weinheim: VCH. doi:10.1002/bbpc.19920960429. ISBN   3-527-28389-7.