Vous êtes sur la page 1sur 44

Ligand Chemistry

Sandeep Badarla

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information. PDF generated at: Sun, 05 Aug 2012 04:56:58 UTC

Contents
Articles
Ligand Crystal field theory Denticity Chelation Hapticity Trans-spanning ligand Linkage isomerism Bridging ligand Metalligand multiple bond Non-innocent ligand Chiral ligand Ligand dependent pathway Ligand field theory 1 9 14 16 20 23 24 25 27 29 34 37 37

References
Article Sources and Contributors Image Sources, Licenses and Contributors 40 41

Article Licenses
License 42

Ligand

Ligand
In coordination chemistry, a ligand is an ion or molecule (see also: functional group) that binds to a central metal atom to form a coordination complex. The bonding between metal and ligand generally involves formal donation of one or more of the ligand's electron deficient pairs. 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 involving Lewis acidic "ligands."[1][2] Metals and metalloids are bound to ligands in virtually all circumstances, although gaseous "naked" metal ions can be generated Cobalt complex [HCo(CO)4] with five ligands in high vacuum. Ligands in a complex dictate the reactivity of the central atom, including ligand substitution rates, the reactivity of the ligands themselves, and redox. Ligand selection is a critical consideration in many practical areas, including bioinorganic and medicinal chemistry, homogeneous catalysis, and environmental chemistry. Ligands are classified in many ways: their charge, their size (bulk), the identity of the coordinating atom(s), and the number of electrons donated to the metal (denticity or hapticity). The size of a ligand is indicated by its cone angle.

History
The composition of coordination complexes have been known since the early 1800s, e.g. Prussian blue and copper vitriol. The key breakthrough occurred when Alfred Werner reconciled formulas and isomers. He showed, among other things, that the formulas of many cobalt(III) and chromium(III) compounds can be understood if the metal has six ligands in an octahedral geometry. The first to use the term "ligand" were Alfred Stock and Carl Somiesky, in relation to silicon chemistry. The theory allows one to understand the difference between coordinated and ionic chloride in the cobalt ammine chlorides and to explain many of the previously inexplicable isomers. He resolved the first coordination complex called hexol into optical isomers, overthrowing the theory that chirality was necessarily associated with carbon compounds.[3][4]

Strong field and weak field ligands


In general, ligands are viewed as electron donors and the metals as electron acceptors(citation. metals are electron donors). Bonding is often described using the formalisms of molecular orbital theory. The HOMO (Highest Occupied Molecular Orbital) can be mainly of ligands or metal character. Ligands and metal ions can be ordered in many ways; one ranking system focuses on ligand 'hardness' (see also hard/soft acid/base theory). Metal ions preferentially bind certain ligands. In general, 'soft' metal ions prefer weak field ligands, whereas 'hard' metal ions prefer strong field ligands. According to the molecular orbital theory, the HOMO of the ligand should have an energy that overlaps with the LUMO (Lowest Unoccupied Molecular Orbital) of the metal preferential. Metal ions bound to strong-field ligands follow the Aufbau principle, whereas complexes bound to weak-field ligands follow Hund's rule. Binding of the metal with the ligands results in a set of molecular orbitals, where the metal can be identified with a new HOMO and LUMO (the orbitals defining the properties and reactivity of the resulting complex) and a certain ordering of the 5 d-orbitals (which may be filled, or partially filled with electrons). In an octahedral environment, the 5 otherwise degenerate d-orbitals split in sets of 2 and 3 orbitals (for a more in depth explanation, see crystal field theory).

Ligand 3 orbitals of low energy: dxy, dxz and dyz 2 of high energy: dz2 and dx2y2 The energy difference between these 2 sets of d-orbitals is called the splitting parameter, o. The magnitude of o is determined by the field-strength of the ligand: strong field ligands, by definition, increase o more than weak field ligands. Ligands can now be sorted according to the magnitude of o (see the table below). This ordering of ligands is almost invariable for all metal ions and is called spectrochemical series. For complexes with a tetrahedral surrounding, the d-orbitals again split into two sets, but this time in reverse order. 2 orbitals of low energy: dz2 and dx2y2 3 orbitals of high energy: dxy, dxz and dyz The energy difference between these 2 sets of d-orbitals is now called t. The magnitude of t is smaller than for o, because in a tetrahedral complex only 4 ligands influence the d-orbitals, whereas in an octahedral complex the d-orbitals are influenced by 6 ligands. When the coordination number is neither octahedral nor tetrahedral, the splitting becomes correspondingly more complex. For the purposes of ranking ligands, however, the properties of the octahedral complexes and the resulting o has been of primary interest. The arrangement of the d-orbitals on the central atom (as determined by the 'strength' of the ligand), has a strong effect on virtually all the properties of the resulting complexes. E.g. the energy differences in the d-orbitals has a strong effect in the optical absorption spectra of metal complexes. It turns out that valence electrons occupying orbitals with significant 3d-orbital character absorb in the 400-800nm region of the spectrum (UV-visible range). The absorption of light (what we perceive as the color) by these electrons (that is, excitation of electrons from one orbital to another orbital under influence of light) can be correlated to the ground state of the metal complex, which reflects the bonding properties of the ligands. The relative change in (relative) energy of the d-orbitals as a function of the field-strength of the ligands is described in Tanabe-Sugano diagrams. In cases where the ligand has low energy LUMO, such orbitals also participate in the bonding. The metal-ligand bond can be further stabilised by a formal donation of electron density back to the ligand in a process known as back-bonding. In this case a filled, central-atom-based orbital donates density into the LUMO of the (coordinated) ligand. Carbon monoxide is the preeminent example a ligand that engages metals via back-donation. Complementarily, ligands with low-energy filled orbitals of pi-symmetry can serve as pi-donor.

Classification of ligands as L and X


Especially in the area of organometallic chemistry, ligands are classified as L and X (or combinations of the two). The classification scheme - the "CBC Method" for Covalent Bond Classification - was popularized by M.L.H. Green and "is based on the notion that there are three basic types [of ligands]... represented by the symbols L, X, and Z, which correspond respectively to 2-electron, 1-electron and 0-electron neutral ligands."[5][6] L ligands are derived from charge-neutral precursors and are represented by amines, phosphines, CO, N2, and alkenes. X ligands typically are derived from anionic precursors such as chloride but includes ligands where salts of anion do not really exist such as hydride and alkyl. Thus, the complex IrCl(CO)(PPh3)2 is classified as an MXL3 complex, since CO and the two PPh3 ligands are classified as L's. The oxidative addition of H2 to IrCl(CO)(PPh3)2 gives an 18e- ML3X3 product, IrClH2(CO)(PPh3)2. EDTA4- is classified as an L2X4 ligand, as it features four anions and two neutral donor sites. Cp is classified as an L2X ligand.[7]

Metal-EDTA complex, wherein the aminocarboxylate is a hexadentate chelating ligand.

Ligand

Polydentate and polyhapto ligand motifs and nomenclature


Denticity
Denticity (represented by ) refers to the number of times a ligand bonds to a metal Cobalt(III) complex containing through non-contiguous donor sites. Many ligands are capable of binding metal ions six ammonia ligands, which are monodentate. The chloride is not through multiple sites, usually because the ligands have lone pairs on more than one a ligand. atom. Ligands that bind via more than one atom are often termed chelating. A ligand that binds through two sites is classified as bidentate, and three sites as tridentate. The "bite angle" refers to the angle between the two bonds of a bidentate chelate. Chelating ligands are commonly formed by linking donor groups via organic linkers. A classic bidentate ligand is ethylenediamine, which is derived by the linking of two ammonia groups with an ethylene (-CH2CH2-) linker. A classic example of a polydentate ligand is the hexadentate chelating agent EDTA, which is able to bond through six sites, completely surrounding some metals. The number of times a polydentate ligand bind to a metal centre is symbolized with "n", where "n" indicates the number sites by which a ligand attaches to a metal. EDTA4, when it is hexidentate, binds as a 6-ligand, the amines and the carboxylate oxygen atoms are not contiguous. In practice, the n value of a ligand is not indicated explicitly but rather assumed. The binding affinity of a chelating system depends on the chelating angle or bite angle. Complexes of polydentate ligands are called chelate complexes. They tend to be more stable than complexes derived from monodentate ligands. This enhanced stability, the chelate effect, is usually attributed to effects of entropy, which favors the displacement of many ligands by one polydentate ligand. When the chelating ligand forms a large ring that at least partially surrounds the central atom and bonds to it, leaving the central atom at the centre of a large ring. The more rigid and the higher its denticity, the more inert will be the macrocyclic complex. Heme is a good example: the iron atom is at the centre of a porphyrin macrocycle, being bound to four nitrogen atoms of the tetrapyrrole macrocycle. The very stable dimethylglyoximate complex of nickel is a synthetic macrocycle derived from the anion of dimethylglyoxime.

Hapticity
Hapticity (represented by ) refers to the number of contiguous atoms that comprise a donor site and attach to a metal center. Butadiene forms both 2 and 4 complexes depending on the number of carbon atoms that are bonded to the metal.[7]

Ligand motifs
Outer-sphere ligands
In coordination chemistry, the ligands that are directly bonded to the metal (that is, share electrons), form part of the first coordination sphere and are sometimes called "inner sphere" ligands. "Outer-sphere" ligands are not directly attached to the metal, but are bonded, generally weakly, to the first coordination shell, affecting the inner sphere in subtle ways. The complex of the metal with the inner sphere ligands is then called a coordination complex, which can be neutral, cationic, or anionic. The complex, along with its counterions (if required), is called a coordination compound.

Ligand

Trans-spanning ligands
Trans-spanning ligands are bidentate ligands that can span coordination positions on opposite sides of a coordination complex.[8]

Ambidentate ligand
Unlike polydentate ligands, ambidentate ligands can attach to the central atom in two places but not both. A good example of this is thiocyanate, SCN, which can attach at either the sulfur atom or the nitrogen atom. Such compounds give rise to linkage isomerism. Polyfunctional ligands, see especially proteins, can bond to a metal center through different ligand atoms to form various isomers.

Bridging ligand
A bridging ligand links two or more metal center. Virtually all inorganic solids with simple formulas are coordination polymers, consisting of metal centres linked by bridging ligands. This group of materials includes all anhydrous binary metal halides and pseudohalides. Bridging ligands also persist in solution. Polyatomic ligands such as carbonate are ambidentate and thus are found to often bind to two or three metals simultaneously. Atoms that bridge metals are sometimes indicated with the prefix "" (mu). Most inorganic solids, are polymers by virtue of the presence of multiple bridging ligands.

Metalligand multiple bond


Metal ligand multiple bonds some ligands can bond to a metal center through the same atom but with a different number of lone pairs. The bond order of the metal ligand bond can be in part distinguished through the metal ligand bond angle (M-X-R). This bond angle is often referred to as being linear or bent with further discussion concerning the degree to which the angle is bent. For example, an imido ligand in the ionic form has three lone pairs. One lone pair is used as a sigma X donor, the other two lone pairs are available as L type pi donors. If both lone pairs are used in pi bonds then the M-N-R geometry is linear. However, if one or both these lone pairs is non-bonding then the M-N-R bond is bent and the extent of the bend speaks to how much pi bonding there may be. 1-Nitric oxide can coordinate to a metal center in linear or bent manner.

Specialized ligand types


Non-innocent ligand
Non-innocent ligands bond with metals in such a manner that the distribution of electron density between the metal center and ligand is unclear. Describing the bonding of noninnocent ligands often involves writing multiple resonance forms that have partial contributions to the overall state.

Bulky ligands
Bulky ligands are used to control the steric properties of a metal center. They are used for many reasons, both practical and academic. On the practical side, they influence the selectivity of metal catalysts, e.g. in hydroformylation. Of academic interest, bulky ligands stabilize unusual coordination sites, e.g. reactive coligands or low coordination numbers. Often bulky ligands are employed to simulate the steric protection afforded by proteins to metal-containing active sites. Of course excessive steric bulk can prevent the coordination of certain ligands.

Ligand

Chiral ligands
Chiral ligands are useful for inducing asymmetry within the coordination sphere. Often the ligand is employed as an optically pure group. In some cases, e.g. secondary amines, the asymmetry arises upon coordination. Chiral ligands are essential components of asymmetric homogeneous catalysis.

Common ligands
See nomenclature. Virtually every molecule and every ion can serve as a ligand for (or "coordinate to") metals. Monodentate ligands include virtually all anions and all simple Lewis bases. Thus, the halides and pseudohalides are important anionic ligands whereas ammonia, carbon monoxide, and water are particularly common charge-neutral ligands. Simple organic species are also very common, be they anionic (RO and RCO2) or neutral (R2O, R2S, R3xNHx, and R3P). The steric properties of some ligands are evaluated in terms of their cone angles. Beyond the classical Lewis bases and anions, all unsaturated molecules are also ligands, utilizing their -electrons in forming the coordinate bond. Also, metals can bind to the bonds in for example silanes, hydrocarbons, and dihydrogen (see also: agostic interaction). In complexes of non-innocent ligands, the ligand is bonded to metals via conventional bonds, but the ligand is also redox-active.

Examples of common ligands (by field strength)


In the following table the ligands are sorted by field strength (weak field ligands first):
Ligand formula (bonding atom(s) in bold) I Br S2 Charge Most common denticity Remark(s)

Iodide (iodo) Bromide (bromido) Sulfide (thio or less commonly "bridging thiolate")

monoanionic monodentate monoanionic monodentate dianionic monodentate (M=S), or bidentate bridging (M-S-M') ambidentate (see also isothiocyanate, below) also found bridging

Thiocyanate (S-thiocyanato)

S-CN Cl O-NO2 N-N2 F O-H [O-C(=O)-C(=O)-O]2 H-O-H O-N-O N=C=S CH3CN

monoanionic monodentate

Chloride (chlorido) Nitrate (nitrato) Azide (azido) Fluoride (fluoro) Hydroxide (hydroxo) Oxalate (oxalato) Water (aqua) Nitrite (nitrito) Isothiocyanate (isothiocyanato)

monoanionic monodentate monoanionic monodentate monoanionic monodentate monoanionic monodentate monoanionic monodentate dianionic neutral bidentate monodentate

often found as a bridging ligand

monodentate ambidentate (see also nitro) ambidentate (see also thiocyanate, above)

monoanionic monodentate monoanionic monodentate

Acetonitrile (acetonitrilo)

neutral

monodentate

Ligand

6
C5H5N NH3 en bipy neutral neutral monodentate monodentate

Pyridine Ammonia (ammine or less commonly "ammino") Ethylenediamine 2,2'-Bipyridine

neutral neutral

bidentate bidentate easily reduced to its (radical) anion or even to its dianion

1,10-Phenanthroline Nitrite (nitro) Triphenylphosphine Cyanide (cyano)

phen N-O2 PPh3 CN

neutral

bidentate ambidentate (see also nitrito)

monoanionic monodentate neutral monodentate

monoanionic monodentate

can bridge between metals (both metals bound to C, or one to C and one to N) can bridge between metals (both metals bound to C)

Carbon monoxide (carbonyl)

CO

neutral

monodentate

Note: The entries in the table are sorted by field strength, binding through the stated atom (i.e. as a terminal ligand), the 'strength' of the ligand changes when the ligand binds in an alternative binding mode (e.g. when it bridges between metals) or when the conformation of the ligand gets distorted (e.g. a linear ligand that is forced through steric interactions to bind in a non-linear fashion).

Other general encountered ligands (alphabetical)


In this table other common ligands are listed in alphabetical order.
Ligand formula (bonding atom(s) in bold) Charge Most common denticity Remark(s)

Acetylacetonate (Acac)

CH3-C(O)-CH2-C(O)-CH3

monoanionic bidentate

In general bidentate, bound through both oxygens, but sometimes bound through the central carbon only, see also analogous ketimine analogues compounds with a C-C double bond and other arenes

Alkenes

R2C=CR2 C6H6 Ph2PC2H4PPh2 C25H22P2

neutral

Benzene 1,2-Bis(diphenylphosphino)ethane (dppe) 1,1-Bis(diphenylphosphino)methane (dppm)

neutral neutral neutral bidentate

Can bond to 2 metal atoms at once, forming dimers tetradentate

Corroles Crown ethers neutral

primarily for alkali and alkaline earth metal cations hexadentate primarily for alkali and alkaline earth metal cations

2,2,2-crypt

Ligand

7
neutral [C5H5] monoanionic Although monoanionic, by the nature of its occupied MO's, it is capable of acting as a tridentate ligand. tridentate related to TACN, but not constrained to facial complexation

Cryptates Cyclopentadienyl (Cp)

Diethylenetriamine (dien)

C4H13N3

neutral

Dimethylglyoximate (dmgH) Ethylenediaminetetraacetate (EDTA) (HOOC-CH2)2N-(CH2)2-N(CH2-COOH)2

monoanionic tetra-anionic hexadentate actual ligand is the tetra-anion

Ethylenediaminetriacetate

trianionic

pentadentate actual ligand is the trianion

Ethyleneglycol-bis(oxyethylenenitrilo)-tetraacetate (HOOC-CH2)2N-(CH2)2-O-(CH2)2-O-(CH2)2-N(CH2-COOH)2 tetra-anionic octodentate (EGTA) glycinate (Glycinato) monoanionic bidentate other -amino acid anions are comparable (but chiral) macrocyclic ligand bent (1e) and linear (3e) bonding mode monodentate sometimes bridging ditopic tridentate monoanionic monodentate ambidentate neutral tridentate meridional bonding only macrocyclic ligand see also the N,N',N"-trimethylated analogue sometimes bridging

Heme Nitrosyl NO+ O N2C4H4

dianionic cationic

tetradentate

Oxo Pyrazine Scorpionate ligand Sulfite 2,2',5',2-Terpyridine (terpy)

dianion neutral

Triazacyclononane (tacn)

(C2H4)3(NR)3

neutral

tridentate

Tricyclohexylphosphine Triethylenetetramine (trien) Trimethylphosphine Tri(o-tolyl)phosphine Tris(2-aminoethyl)amine (tren) Tris(2-diphenylphosphineethyl)amine (np3) Terpyridine Tropylium Carbon dioxide

(C6H11)3P or (PCy3)

neutral neutral

monodentate tetradentate monodentate monodentate tetradentate tetradentate tridentate

PMe3 P(o-tolyl)3 (NH2CH2CH2)3N

neutral neutral neutral neutral

C15H11N3 C7H7+ CO2

neutral cationic

see transition metal carbon dioxide complex

Ligand

Ligand exchange
Ligand exchange (also ligand substitution) is a type of chemical reaction in which one ligand in a chemical compound is replaced by another ligand. One type of pathway for substitution is the Ligand dependent pathway. In organometallic chemistry this can take place by associative substitution or by dissociative substitution. Another form of ligand exchange is seen in the nucleophilic abstraction reaction.

Pronunciation
Pronounced lgnd with the first syllable sounding like the word "lithium" or 'lagnd with the first syllable sounding like the word "lie". [9]

References
[1] Cotton, Frank Albert; Geoffrey Wilkinson, Carlos A. Murillo (1999). Advanced Inorganic Chemistry. pp.1355. ISBN0-471-19957-5, 9780471199571. [2] Miessler, Gary L.; Donald Arthur Tarr (1999). Inorganic Chemistry. pp.642. ISBN0-13-841891-8, 9780138418915. [3] Jackson, W. Gregory; Josephine A. McKeon, Silvia Cortez (1 October 2004). "Alfred Werner's Inorganic Counterparts of Racemic and Mesomeric Tartaric Acid: A Milestone Revisited". Inorganic Chemistry 43 (20): 62496254. doi:10.1021/ic040042e. PMID15446870. [4] Bowman-James, Kristin (2005). "Alfred Werner Revisited: The Coordination Chemistry of Anions". Accounts of Chemical Research 38 (8): 671678. doi:10.1021/ar040071t. PMID16104690. [5] Green, M. L. H. (20 September 1995). "A new approach to the formal classification of covalent compounds of the elements". Journal of Organometallic Chemistry 500 (12): 127148. doi:10.1016/0022-328X(95)00508-N. ISSN0022-328X. [6] http:/ / www. columbia. edu/ cu/ chemistry/ groups/ parkin/ mlxz. htm [7] Hartwig, J. F. Organotransition Metal Chemistry, from Bonding to Catalysis; University Science Books: New York, 2010. ISBN 1-891389-53-X [8] von Zelewsky, A. "Stereochemistry of Coordination Compounds" John Wiley: Chichester, 1995. ISBN 047195599. [9] "Ligand - Definition and more from the Free Merriam-Webster Dictionary" (http:/ / www. merriam-webster. com/ dictionary/ ligand?show=0& t=1303577063). Merriam Webster. . Retrieved 23 April 2011.

Crystal field theory

Crystal field theory


Crystal field theory (CFT) is a model that describes the breaking of degeneracies of electronic orbital states, usually d or f orbitals, due to a static electric field produced by a surrounding charge distribution (anion neighbors). This theory has been used to describe various spectroscopies of transition metal coordination complexes, in particular optical spectra (colours). CFT successfully accounts for some magnetic properties, colours, 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[1] 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.

Overview of crystal field theory analysis


According to CFT, the interaction between a transition metal and ligands arises from the attraction between the positively charged metal cation and negative charge on the non-bonding electrons of the ligand. The theory is developed by considering energy changes of the five degenerate d-orbitals upon being surrounded by an array of point charges consisting of the ligands. As a ligand approaches the metal ion, the electrons from the ligand will be closer to some of the d-orbitals and farther away from others causing a loss of degeneracy. The electrons in the d-orbitals and those in the ligand repel each other due to repulsion between like charges. Thus the d-electrons closer to the ligands will have a higher energy than those further away which results in the d-orbitals splitting in energy. This splitting is affected by the following factors: the nature of the metal ion. the metal's oxidation state. A higher oxidation state leads to a larger splitting. the arrangement of the ligands around the metal ion. the nature of the ligands surrounding the metal ion. The stronger the effect of the ligands then the greater the difference between the high and low energy d groups.

The most common type of complex is octahedral; here six ligands form an octahedron around the metal ion. In octahedral symmetry the d-orbitals split into two sets with an energy difference, oct (the crystal-field splitting parameter) where the dxy, dxz and dyz orbitals will be lower in energy than the dz2 and dx2-y2, which will have higher energy, because the former group is farther from the ligands than the latter and therefore experience less repulsion. The three lower-energy orbitals are collectively referred to as t2g, and the two higher-energy orbitals as eg. (These labels are based on the theory of molecular symmetry). Typical orbital energy diagrams are given below in the section High-spin and low-spin. Tetrahedral complexes are the second most common type; here four ligands form a tetrahedron around the metal ion. In a tetrahedral crystal field splitting the d-orbitals again split into two groups, with an energy difference of tet where the lower energy orbitals will be dz2 and dx2-y2, and the higher energy orbitals will be dxy, dxz and dyz opposite to the octahedral case. Furthermore, since the ligand electrons in tetrahedral symmetry are not oriented directly towards the d-orbitals, the energy splitting will be lower than in the octahedral case. Square planar and other complex geometries can also be described by CFT. The size of the gap between the two or more sets of orbitals depends on several factors, including the ligands and geometry of the complex. Some ligands always produce a small value of , while others always give a large splitting. The reasons behind this can be explained by ligand field theory. The spectrochemical series is an empirically-derived list of ligands ordered by the size of the splitting that they produce (small to large ; see also this table): I < Br < S2 < SCN < Cl < NO3 < N3 < F < OH < C2O42 < H2O < NCS < CH3CN < py < NH3 < en < 2,2'-bipyridine < phen < NO2 < PPh3 < CN < CO

Crystal field theory It is useful to note that the ligands producing the most splitting are those that can engage in metal to ligand back-bonding. The oxidation state of the metal also contributes to the size of between the high and low energy levels. As the oxidation state increases for a given metal, the magnitude of increases. A V3+ complex will have a larger than a V2+ complex for a given set of ligands, as the difference in charge density allows the ligands to be closer to a V3+ ion than to a V2+ ion. The smaller distance between the ligand and the metal ion results in a larger , because the ligand and metal electrons are closer together and therefore repel more.

10

High-spin and low-spin


Ligands which cause a large splitting of the d-orbitals are referred to as strong-field ligands, such as CN and CO from the spectrochemical series. In complexes with these ligands, it is unfavourable to put electrons into the high energy orbitals. Therefore, the lower energy orbitals are completely filled before population of the upper sets starts according to the Aufbau principle. Complexes such as this are called "low spin". For Low Spin [Fe(NO2)6]3 crystal field diagram example, NO2 is a strong-field ligand and produces a large . The octahedral ion [Fe(NO2)6]3, which has 5 d-electrons, would have the octahedral splitting diagram shown at right with all five electrons in the t2g level. Conversely, ligands (like I and Br) which cause a small splitting of the d-orbitals are referred to as weak-field ligands. In this case, it is easier to put electrons into the higher energy set of orbitals than it is to put two into the same low-energy orbital, because two electrons in the same orbital repel each other. So, one electron is put into each of the five d-orbitals High Spin [FeBr6]3 crystal field diagram before any pairing occurs in accord with Hund's rule and "high spin" complexes are formed. For example, Br is a weak-field ligand and produces a small oct. So, the ion [FeBr6]3, again with five d-electrons, would have an octahedral splitting diagram where all five orbitals are singly occupied. 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 . As noted above, eg refers to the dz2 and dx2-y2 which are higher in energy than the t2g in octahedral complexes. If the energy required to pair two electrons is greater than the energy cost of placing an electron in an eg, , high spin splitting occurs. The crystal field splitting energy for tetrahedral metal complexes (four ligands) is referred to as tet, and is roughly equal to 4/9oct (for the same metal and same ligands). Therefore, the energy required to pair two electrons is typically higher than the energy required for placing electrons in the higher energy orbitals. Thus, tetrahedral complexes are usually high-spin. The use of these splitting diagrams can aid in the prediction of the magnetic properties of coordination compounds. A compound that has unpaired electrons in its splitting diagram will be paramagnetic and will be attracted by magnetic fields, while a compound that lacks unpaired electrons in its splitting diagram will be diamagnetic and will be weakly repelled by a magnetic field.

Crystal field theory

11

Crystal field stabilization energy


The crystal field stabilization energy (CFSE) is the stability that results from placing a transition metal ion in the crystal field generated by a set of ligands. It arises due to the fact that when the d-orbitals are split in a ligand field (as described above), some of them become lower in energy than before with respect to a spherical field known as the barycenter in which all five d-orbitals are degenerate. For example, in an octahedral case, the t2g set becomes lower in energy than the orbitals in the barycenter. As a result of this, if there are any electrons occupying these orbitals, the metal ion is more stable in the ligand field relative to the barycenter by an amount known as the CFSE. Conversely, the eg orbitals (in the octahedral case) are higher in energy than in the barycenter, so putting electrons in these reduces the amount of CFSE. If the splitting of the d-orbitals in an octahedral field is oct, the three t2g orbitals are stabilized relative to the barycenter by 2/5 oct, and the eg orbitals are destabilized by 3/5 oct. As examples, consider the two d5 Octahedral crystal field stabilization energy configurations shown further up the page. The low-spin (top) example has five electrons in the t2g orbitals, so the total CFSE is 5 x 2/5 oct = 2oct. In the high-spin (lower) example, the CFSE is (3 x 2/5 oct) - (2 x 3/5 oct) = 0 - in this case, the stabilization generated by the electrons in the lower orbitals is canceled out by the destabilizing effect of the electrons in the upper orbitals. Crystal Field stabilization is applicable to transition-metal complexes of all geometries. Indeed, the reason that many d8 complexes are square-planar is the very large amount of crystal field stabilization that this geometry produces with this number of electrons.

Explaining the colors of transition metal complexes


The bright colors exhibited by many coordination compounds can be explained by Crystal Field Theory. If the d-orbitals of such a complex have been split into two sets as described above, when the molecule absorbs a photon of visible light one or more electrons may momentarily jump from the lower energy d-orbitals to the higher energy ones to transiently create an excited state atom. The difference in energy between the atom in the ground state and in the excited state is equal to the energy of the absorbed photon, and related inversely to the wavelength of the light. Because only certain wavelengths () of light are absorbed - those matching exactly the energy difference - the compounds appears the appropriate complementary color. As explained above, because different ligands generate crystal fields of different strengths, different colors can be seen. For a given metal ion, weaker field ligands create a complex with a smaller , which will absorb light of longer and thus lower frequency . Conversely, stronger field ligands create a larger , absorb light of shorter , and thus higher . It is, though, rarely the case that the energy of the photon absorbed corresponds exactly to the size of the gap ; there are other things (such as electron-electron repulsion and Jahn-Teller effects) that also affect the energy difference between the ground and excited states.

Crystal field theory

12

Which colors are exhibited?


This color wheel demonstrates which color a compound will appear if it only has one absorption in the visible spectrum. For example, if the compound absorbs red light, it will appear green. absorbed versus color observed 400nm Violet absorbed, Green-yellow observed ( 560nm) 450nm Blue absorbed, Yellow observed ( 600nm) 490nm Blue-green absorbed, Red observed ( 620nm) 570nm Yellow-green absorbed, Violet observed ( 410nm) 580nm Yellow absorbed, Dark blue observed ( 430nm) 600nm Orange absorbed, Blue observed ( 450nm) 650nm Red absorbed, Green observed ( 520nm)

Color wheel

Crystal field splitting diagrams


Crystal field splitting diagrams (-acceptor ligands)
Octahedral Pentagonal bipyramidal Square antiprismatic

Square planar

Square pyramidal

Tetrahedral

Trigonal bipyramidal

Crystal field theory

13

Gallery

Octahedral

Pentagonal bipyramidal

Square planar

Square pyramidal

Tetrahedral

Trigonal bipyramidal

Pentagonal pyramidal

References
[1] Van Vleck, J. (1932). "Theory of the Variations in Paramagnetic Anisotropy Among Different Salts of the Iron Group". Physical Review 41: 208. Bibcode1932PhRv...41..208V. doi:10.1103/PhysRev.41.208.

Zumdahl, Steven S (2005). Chemical Principles (5th ed.). Houghton Mifflin Company. pp.550551, 957964. ISBN0-669-39321-5. Silberberg, Martin S (2006). Chemistry: The Molecular Nature of Matter and Change (4th ed.). New York: McGraw Hill Company. pp.10281034. ISBN0-8151-8505-7. Shriver, D. F.; Atkins, P. W. (2001). Inorganic Chemistry (4th ed.). Oxford University Press. pp.227236. ISBN0-8412-3849-9. Housecroft, C. E.; Sharpe, A. G. (2004). Inorganic Chemistry (2nd ed.). Prentice Hall. ISBN978-0130399137. Miessler, G. L.; Tarr, D. A. (2003). Inorganic Chemistry (3rd ed.). Pearson Prentice Hall. ISBN0-13-035471-6.

Denticity

14

Denticity
Denticity refers to the number of atoms in a single ligand that bind to a central atom in a coordination complex.[1][2] In many cases, only one atom in the ligand binds to the metal, so the denticity equals one, and the ligand is said to be monodentate (sometimes called unidentate). Ligands with more than one bonded atom are called polydentate or multidentate. The word denticity is derived from dentis, the Latin word for tooth. The ligand is thought of as biting the metal at one or more linkage points. Denticity is distinguished from hapticity, in which electrons of a bond or conjugated series of bonds are linked to the central metal without the metal-ligand bond being localized to a single ligand atom.

Classes of denticity

Polydentate ligands are chelating agents[3] and classified by their denticity. Some atoms cannot form the maximum possible number of bonds a ligand could make. In that case one or more binding sites of the ligand are unused. Such sites can be used to form a bond with another chemical species. Bidentate (also called didentate) ligands bind with two atoms, an example being ethylenediamine. Tridentate ligands bind with three atoms, an example being terpyridine. Tridentate ligands usually bind via two kinds of connectivity, called "mer" and "fac." Cyclic tridentate ligands such as TACN and 9-ane-S3 bind in a facial manner. Tetradentate ligands bind with four atoms, an example being triethylenetetramine (abbreviated trien). Tetradentate ligands bind via three connectivities depending on their topology and the geometry of the metal center. For Structure of the pharmaceutical Oxaliplatin, which features octahedral metals, the linear tetradentate trien can bind via two different bidentate ligands. three geometries. Tripodal tetradentate ligands, e.g. tris(2-aminoethyl)amine, are more constrained, and on octahedra leave two cis sites. Many naturally occurring macrocyclic ligands are tetradentative, an example being the porphyrin in heme. Pentadentate ligands bind with five atoms, an example being ethylenediaminetriacetic acid. Hexadentate ligands bind with six atoms, an example being EDTA (although it can bind in a tetradentate manner). Ligands of denticity greater than 6 are well known. The ligands 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetate (DOTA) and diethylene triamine pentaacetate (DTPA) are octadentate. They are particularly useful for binding lanthanide ions, which typically have coordination numbers greater than 6.

Atom with monodentate ligands

Denticity

15

Relationship between "linear" bi-, tri- and tetradentate ligands (red) bound to an octahedral metal center. The structures marked with * are chiral owing to the backbone of the tetradentate ligand.

Stability constants
In general, the stability of a metal complex correlates with the denticity of the ligands. The stability of a complex is represented quantitatively in the form of Stability constants. Hexadentate ligands tend to bind metal ions more strongly than ligands of lower denticity.

External links
EDTA chelation lecture notes. [4] 2.4MB PDF - Slide 3 on denticity

References
[1] [2] [3] [4] IUPAC Gold Book denticity (http:/ / goldbook. iupac. org/ D01594. html) von Zelewsky, A. "Stereochemistry of Coordination Compounds" John Wiley: Chichester, 1995. ISBN 047195599. IUPAC Gold Book chelation (http:/ / goldbook. iupac. org/ C01012. html) http:/ / people. chem. byu. edu/ dvd/ chem223/ lecture_notes/ 11_EDTA_Chelation. pdf/ at_download/ file

Chelation

16

Chelation
Chelation is the formation or presence of two or more separate coordinate bonds between a polydentate (multiple bonded) ligand and a single central atom.[1] Usually these ligands are organic compounds, and are called chelants, chelators, chelating agents, or sequestering agents. The ligand forms a chelate complex with the substrate. Chelate complexes are contrasted with coordination complexes composed of monodentate ligands, which form only one bond with the central atom. Chelants, according to ASTM-A-380, are "chemicals that form soluble, complex molecules with certain metal ions, inactivating the ions so that they cannot normally react with other elements or ions to produce precipitates or scale." The word chelation is derived from Greek , chel, meaning claw; the ligands lie around the central atom like the claws of a lobster.[2]
Metal-EDTA chelate

The chelate effect


The chelate effect describes the enhanced affinity of chelating ligands for a metal ion compared to the affinity of a collection of similar nonchelating (monodentate) ligands for the same metal. Consider the two equilibria, in aqueous solution, between the copper(II) ion, Cu2+ and ethylenediamine (en) on the one hand and methylamine, MeNH2 on the other. Cu2+ + en [Cu(en)]2+ (1) [Cu(MeNH2)2]2+ (2)
Ethylenediamine ligand, binding to a central

Cu2+ + 2 MeNH2

In (1) the bidentate ligand ethylene diamine forms a chelate complex metal ion with two bonds with the copper ion. Chelation results in the formation of a fivemembered ring. In (2) the bidentate ligand is replaced by two monodentate methylamine ligands of approximately the same donor power, meaning that the enthalpy of formation of CuN bonds is approximately the same in the two reactions. Under conditions of equal copper concentrations and when the concentration of methylamine is twice the concentration of ethylenediamine, the concentration of the complex (1) will be greater than the concentration of the complex (2). Cu2+ complexes with methylamine (left) and ethylenediamine (right) The effect increases with the number of chelate rings so the concentration of the EDTA complex, which has six chelate rings, is much much higher than a corresponding complex with two monodentate nitrogen donor ligands and four monodentate carboxylate ligands. Thus, the phenomenon of the chelate effect is a firmly established empirical fact.

Chelation The thermodynamic approach to explaining the chelate effect considers the equilibrium constant for the reaction: the larger the equilibrium constant, the higher the concentration of the complex. [Cu(en)] =11[Cu][en] [Cu(MeNH2)2]= 12[Cu][MeNH2]2 Electrical charges have been omitted for simplicity of notation. The square brackets indicate concentration, and the subscripts to the stability constants, , indicate the stoichiometry of the complex. When the analytical concentration of methylamine is twice that of ethylenediamine and the concentration of copper is the same in both reactions, the concentration [Cu(en)] is much higher than the concentration [Cu(MeNH2)2] because 11 >> 12. An equilibrium constant, K, is related to the standard Gibbs free energy, G G = RT ln K = H TS is the standard enthalpy change of the reaction by

17

where R is the gas constant and T is the temperature in kelvins. H and S is the standard entropy change.

It has already been posited that the enthalpy term should be approximately the same for the two reactions. Therefore the difference between the two stability constants is due to the entropy term. In equation (1) there are two particles on the left and one on the right, whereas in equation (2) there are three particles on the left and one on the right. This means that less entropy of disorder is lost when the chelate complex is formed than when the complex with monodentate ligands is formed. This is one of the factors contributing to the entropy difference. Other factors include solvation changes and ring formation. Some experimental data to illustrate the effect are shown in the following table.[3]
Equilibrium Cd2+ + 4 MeNH2 Cd2+ + 2 en Cd(MeNH2)42+ log G 6.55 -37.4 H -57.3 -56.48 /kJ mol1 TS 19.9 -4.19 /kJ mol1

Cd(en)22+

10.62 -60.67

These data show that the standard enthalpy changes are indeed approximately equal for the two reactions and that the main reason for the greater stability of the chelate complex is due to the entropy term, which is much less unfavourable, indeed, it is favourable in this instance. In general it is difficult to account precisely for thermodynamic values in terms of changes in solution at the molecular level, but it is clear that the chelate effect is predominantly an effect of entropy. Other explanations, including that of Schwarzenbach,[4] are discussed in Greenwood and Earnshaw (loc.cit).

In nature
Virtually all biochemicals exhibit the ability to dissolve certain metal cations. Thus, proteins, polysaccharides, and polynucleic acids are excellent polydentate ligands for many metal ions. Organic compounds such as the amino acids glutamic acid and histidine, organic diacids such as malate, and polypeptides such as phytochelatin are also typical chelators. In addition to these adventitious chelators, several biomolecules are specifically produced to bind certain metals (see next section).[5][6][7][8]

In biochemistry and microbiology


Virtually all metalloenzymes feature metals that are chelated, usually to peptides or cofactors and prosthetic groups.[8] Such chelating agents include the porphyrin rings in hemoglobin and chlorophyll. Many microbial species produce water-soluble pigments that serve as chelating agents, termed siderophores. For example, species of Pseudomonas are known to secrete pyocyanin and pyoverdin that bind iron. Enterobactin, produced by E. coli, is the strongest chelating agent known.

Chelation

18

In geology
In earth science, chemical weathering is attributed to organic chelating agents, e.g. peptides and sugars, that extract metal ions from minerals and rocks.[9] Most metal complexes in the environment and in nature are bound in some form of chelate ring, e.g. with a humic acid or a protein. Thus, metal chelates are relevant to the mobilization of metals in the soil, the uptake and the accumulation of metals into plants and microorganisms. Selective chelation of heavy metals is relevant to bioremediation, e.g. removal of 137Cs from radioactive waste.[10]

Applications
Chelators are used in producing nutritional supplements, fertilizers, chemical analysis, as water softeners, commercial products such as shampoos and food preservatives, medicine, heavy metal detox, and industrial applications. In 2010, the Asia-Pacific region was the largest outlet, generating about 45% of worldwide demand for chelating agents. The region was followed by Western Europe and North America. The global chelating agents market is expected to reach more than 5 million tonnes in 2018.[11]

Nutritional supplements
In the 1960's, scientists developed the concept of chelating a metal ion prior to feeding the element to the animal. They believed that this would create a neutral compound, protecting the mineral from being complexed with insoluble salts within the stomach, rendering the metal unavailable for absorption. Amino acids, being effective metal binders, were chosen as the prospective ligands, and research was conducted on the metal-amino acid combinations. The research supported that the metal-amino acid chelates were able to enhance mineral absorption. During this period, synthetic chelates were also being developed. An example of such synthetics is ethylenediaminetetraacetic acid (EDTA). These synthetics applied the same concept of chelation and did create chelated compounds; however, these synthetics were too stable and not nutritionally viable. If the mineral was taken from the EDTA ligand, the ligand could not be used by the body and would be expelled. During the expulsion process the EDTA ligand will randomly chelate and strip another mineral from the body.[12] According to the Association of American Feed Control Officials (AAFCO), a metal amino acid chelate is defined as the product resulting from the reaction of a metal ion from a soluble metal salt with a mole ratio of one to three (preferably two) moles of amino acids. The average weight of the hydrolyzed amino acids must be approximately 150 and the resulting molecular weight of the chelate must not exceed 800 Da. Since the early development of these compounds, much more research has been conducted, and has been applied to human nutrition products in a similar manner to the animal nutrition experiments that pioneered the technology. Ferrous bis-glycinate is an example of one of these compounds that has been developed for human nutrition. [13]

Fertilizers
Many mineral deficiencies can occur in plants, such as iron chlorosis, which can reduce the nutritional benefits of crops and eventually result in plant death. Mineral chelates have been used to alleviate the mineral deficiencies of affected crops through liquid foliar applications. These fertilizers are also used to prevent deficiencies from occurring and improving the overall health of the plants. [14]

Heavy metal detoxification


Chelation therapy is the use of chelating agents to detoxify poisonous metal agents such as mercury, arsenic, and lead by converting them to a chemically inert form that can be excreted without further interaction with the body, and was approved by the U.S. Food and Drug Administration in 1991. In alternative medicine, chelation is used as a treatment for autism, although this practice is controversial due to the absence of scientific plausibility, lack of FDA

Chelation approval, and its potentially deadly side-effects.[15] Although they can be beneficial in cases of heavy metal poisoning, chelating agents can also be dangerous. Use of disodium EDTA instead of calcium EDTA has resulted in fatalities due to hypocalcemia.[16]

19

Other medical applications


Antibiotic drugs of the tetracycline family are chelators of Ca2+ and Mg2+ ions. EDTA is also used in root canal treatment as a way to irrigate the canal. EDTA softens the dentin facilitating access to the entire canal length and to remove the smear layer formed during instrumentation. Chelate complexes of gadolinium are often used as contrast agents in MRI scans.

Chemical applications
Homogeneous catalysts are often chelated complexes. A typical example is the ruthenium(II) chloride chelated with BINAP (a bidentate phosphine) used in e.g. Noyori asymmetric hydrogenation and asymmetric isomerization. The latter has the practical use of manufacture of synthetic ()-menthol. Citric acid is used to soften water in soaps and laundry detergents. A common synthetic chelator is EDTA. Phosphonates are also well known chelating agents. Chelators are used in water treatment programs and specifically in steam engineering, e.g., boiler water treatment system: Chelant Water Treatment system. Products such as Bio-Rust and Evapo-Rust are chelating agents sold for the removal of rust from iron and steel.

References
[1] IUPAC definition of chelation. (http:/ / goldbook. iupac. org/ C01012. html) [2] The term chelate was first applied in 1920 by Sir Gilbert T. Morgan and H. D. K. Drew, who stated: "The adjective chelate, derived from the great claw or chele (Greek) of the lobster or other crustaceans, is suggested for the caliperlike groups which function as two associating units and fasten to the central atom so as to produce heterocyclic rings." Morgan, Gilbert T.; Drew, Harry D. K. (1920). "CLXII.Researches on residual affinity and co-ordination. Part II. Acetylacetones of selenium and tellurium". J. Chem. Soc., Trans. 117: 1456. doi:10.1039/CT9201701456. (nonfree access) [3] Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (http:/ / www. amazon. com/ Chemistry-Elements-Second-Edition-Earnshaw/ dp/ 0750633654) (2nd ed.). ButterworthHeinemann. ISBN0080379419. . p 910 [4] Schwarzenbach, G (1952). "Der Chelateffekt". Helv. Chim. Acta 35 (7): 23442359. doi:10.1002/hlca.19520350721. [5] U Krmer, J D Cotter-Howells, J M Charnock, A H J M Baker, J A C Smith (1996). "Free histidine as a metal chelator in plants that accumulate nickel". Nature 379 (6566): 635638. doi:10.1038/379635a0. [6] Jurandir Vieira Magalhaes (2006). "Aluminum tolerance genes are conserved between monocots and dicots". Proc Natl Acad Sci USA 103 (26): 97499750. doi:10.1073/pnas.0603957103. PMC1502523. PMID16785425. [7] Suk-Bong Ha, Aaron P. Smith, Ross Howden, Wendy M. Dietrich, Sarah Bugg, Matthew J. O'Connell, Peter B. Goldsbrough, and Christopher S. Cobbett (1999). "Phytochelatin synthase genes from Arabidopsis and the yeast Schizosaccharomyces pombe" (http:/ / www. plantcell. org/ cgi/ content/ full/ 11/ 6/ 1153?ck=nck). Plant Cell 11 (6): 11531164. doi:10.1105/tpc.11.6.1153. PMC144235. PMID10368185. . [8] S. J. Lippard, J. M. Berg Principles of Bioinorganic Chemistry University Science Books: Mill Valley, CA; 1994. ISBN 0-935702-73-3. [9] Dr. Michael Pidwirny, University of British Columbia Okanagan, http:/ / www. physicalgeography. net/ fundamentals/ 10r. html [10] Prasad (ed). Metals in the Environment. University of Hyderabad. Dekker, New York, 2001 [11] Market Study on Chelating Agents by Ceresana Research (http:/ / www. ceresana. com/ en/ market-studies/ additives/ chelating-agents). [12] Ashmead, H. DeWayne (1993). The Roles of Amino Acid Chelates in Animal Nutrition. Westwood: Noyes Publications. [13] Albion Laboratories, Inc. "Albion Ferrochel Website" (http:/ / www. albionferrochel. com). . Retrieved July 12, 2011. [14] Ashmead, et al., [ed], H. DeWayne (1986). Foliar Feeding of Plants with Amino Acid Chelates. Park Ridge: Noyes Publications. [15] Doja A, Roberts W (2006). "Immunizations and autism: a review of the literature". Can J Neurol Sci 33 (4): 34146. PMID17168158. [16] U.S. Centers for Disease Control, "Deaths Associated with Hypocalcemia from Chelation Therapy" (March 3, 2006), http:/ / www. cdc. gov/ mmwr/ preview/ mmwrhtml/ mm5508a3. htm

Hapticity

20

Hapticity
The term hapticity is used to describe how a group of contiguous atoms of a ligand are coordinated to a central atom. Hapticity of a ligand is indicated by the Greek character 'eta', . A superscripted number following the denotes the number of contiguous atoms of the ligand that are bound to the metal. In general the -notation is only used when there is more than one atom coordinated (otherwise the -notation is used, see also hapticity vs. denticity).

History
Ferrocene contains two The need for additional nomenclature for organometallic compounds became apparent in 5-cyclopentadienyl the mid-1950s when Dunitz, Orgel, and Rich described the structure of the "sandwich ligands complex" ferrocene by X-ray crystallography[1] where an iron atom is "sandwiched" between two parallel cyclopentadienyl rings. Cotton later proposed the term hapticity derived from the adjectival prefix hapto (from the Greek haptein, to fasten, denoting contact or combination) placed before the name of the olefin,[2] where the Greek letter (eta) is used to denote the number of contiguous atoms of a ligand that bind to a metal center. The term is usually employed to describe ligands containing extended -systems or where agostic bonding is not obvious from the formula.

Historically important compounds where the ligands are described with hapticity
Ferrocene - bis(5-cyclopentadienyl)iron Uranocene - bis(8-1,3,5,7-cyclooctatetraene)uranium W(CO)3(PPri3)2(2-H2 ) - the first compound to be synthesized with a dihydrogen ligand.[3][4] IrCl(CO)[P(C6H5)3]2(2-O2) - the dioxygen derivative which forms reversibly upon oxygenation of Vaska's complex.

Examples
The -notation is encountered in many coordination compounds: Side-on bonding of molecules containing -bonds like H2: W(CO)3(PiPr3)2(2-H2)[3][4] Side-on bonded ligands containing multiple bonded atoms, e.g. ethylene in Zeise's salt or with fullerene, which is bonded through donation of the -bonding electrons: K[PtCl3(2-C2H4)].H2O Related complexes containing bridging -ligands: (-2:2-C2H2)Co2(CO)6 and (Cp*2Sm)2(-2:2- N2)[5] Dioxygen in bis{(trispyrazolylborato)copper(II)}(-2:2-O2), Note that with some bridging ligands, an alternative bridging mode is observed, e.g. 1,1, like in (Me3SiCH2)3V(-N2-1(N),1(N'))V(CH2SiMe3)3 contains a bridging dinitrogen molecule, where the molecule is end-on coordinated to the two metal centers (see hapticity vs. denticity). The bonding of -bonded species can be extended over several atoms, e.g. in allyl, butadiene ligands, but also in cyclopentadienyl or benzene rings can share their electrons. Apparent violations of the 18-electron rule sometimes are explicable in compounds with unusual hapticities:

Hapticity The 18-VE complex (5-C5H5)Fe(1-C5H5)(CO)2 contains one 5 bonded cyclopentadienyl, and one 1 bonded cyclopentadienyl. Reduction of the 18-VE compound [Ru(6-C6Me6)2]2+ (where both aromatic rings are bonded in an 6-coordination), results in another 18VE compound: [Ru(6-C6Me6)(4-C6Me6)]. Examples of polyhapto coordinated heterocyclic and inorganic rings: Cr(5-C4H4S)(CO)3 contains the sulfur heterocycle thiophene and Cr(6-B3N3Me6)(CO)3 contains a coordinated inorganic ring (B3N3 ring).

21

Electrons donated by "- ligands" vs. hapticity


Ligand Electrons Electrons contributed contributed (neutral counting) (ionic counting) 1 3 2 4

1 Allyl 3-Allyl cyclopropenyl 3-Allenyl 2-Butadiene 4-Butadiene 1-cyclopentadienyl 3-cyclopentadienyl 5-cyclopentadienyl pentadienyl cyclohexadienyl 2-Benzene 4-Benzene 6-Benzene 7-Cycloheptatrienyl

3 2 4 1 3 5

4 2 4 2 4 6

2 4 6 7

2 4 6 6 10

8-Cyclooctatetraenyl 8

Changes in hapticity
The hapticity of a ligand can change in the course of a reaction.[6] E.g. in a redox reaction:

Hapticity Here one of the 6-benzene rings changes to a 4-benzene. Similarly hapticity can change during a substitution reaction:

22

Here the 5-cyclopentadienyl changes to an 3-cyclopentadienyl, giving room on the metal for an extra 2-electron donating ligand 'L'. Removal of one molecule of CO and again donation of two more electrons by the cyclopentadienyl ligand restores the 5-cyclopentadienyl. The so-called indenyl effect also describes changes in hapticity in a substitution reaction.

Hapticity vs. denticity


Hapticity must be distinguished from denticity. Polydentate ligands coordinate via multiple coordination sites within the ligand. In this case the coordinating atoms are identified using the -notation, as for example seen in coordination of 1,2-bis(diphenylphosphino)ethane (Ph2PCH2CH2PPh2), to NiCl2 as dichloro[ethane-1,2-diylbis(diphenylphosphane)-2P]nickel(II). If the coordinating atoms are contiguous (connected to each other), the -notation is used, as e.g. in titanocene dichloride: 5 [7] dichlorobis( -2,4-cyclopentadien-1-yl)titanium.

Hapticity and fluxionality


Molecules with polyhapto ligands are often "fluxional", also known as stereochemically non-rigid. Two classes of fluxionality are prevalent for organometallic complexes of polyhapto ligands: Case 1, typically: when the hapticity value is less than the number of sp2 carbon atoms. In such situations, the metal will often migrate from carbon to carbon, maintaining the same net hapticity. The 1-C5H5 ligand in (5-C5H5)Fe( 1-C5H5)(CO)2 rearranges rapidly in solution such that Fe binds alternatingly to each carbon atom in the 1-C5H5 ligand. This reaction is degenerate and, in the jargon of organic chemistry, it is an example of a sigmatropic rearrangement. Case 2, typically: complexes containing cyclic polyhapto ligands with maximized hapticity. Such ligands tend to rotate. A famous example is ferrocene, Fe(5-C5H5)2, wherein the Cp rings rotate with a low energy barrier about the principal axis of the molecule that "skewers" each ring (see rotational symmetry). This "ring whizzing" explains, inter alia, why only one isomer can be isolated for Fe(5-C5H4Br)2. In this case, the rotamers are not necessarily degenerate, but the rotational barriers have low energies of activation.

Hapticity

23

References
[1] J. Dunitz, L. Orgel, A. Rich (1956). "The crystal structure of ferrocene". Acta Crystallographica 9 (4): 3735. doi:10.1107/S0365110X56001091. [2] F. A. Cotton (1968). "Proposed nomenclature for olefin-metal and other organometallic complexes". J. Am. Chem. Soc. 90 (22): 62306232. doi:10.1021/ja01024a059. [3] Kubas, Gregory J. (March 1988). "Molecular hydrogen complexes: coordination of a bond to transition metals" (http:/ / pubs. acs. org/ doi/ abs/ 10. 1021/ ar00147a005). Accounts of Chemical Research 21 (3): 120128. doi:10.1021/ar00147a005. . [4] Kubas, Gregory J. (2001). Metal Dihydrogen and -Bond Complexes - Structure, Theory, and Reactivity (http:/ / books. google. com/ ?id=SSn_OQAACAAJ) (1 ed.). New York: Kluwer Academic/Plenum Publishers. ISBN978-0-306-46465-2. LCCN00059283. . [5] D. Sutton (1993). "Organometallic diazo compounds". Chem. Rev. 93 (3): 9951022. doi:10.1021/cr00019a008. [6] Huttner, Gottfried; Lange, Siegfried; Fischer, Ernst O. (1971). "Molecular Structure of Bis(Hexamethylbenzene)-Ruthenium(0)". Angewandte Chemie, International Edition in English 10 (8): 556557. doi:10.1002/anie.197105561. [7] "IR-9.2.4.1 Coordination Compounds: Describing the Constitution of Coordination Compounds: Specifying donor atoms: General" (http:/ / www. iupac. org/ reports/ provisional/ abstract04/ RB-prs310804/ Chap9-3. 04. pdf). Nomenclature of Inorganic Chemistry Recommendations 1990 (the 'Red Book') (http:/ / old. iupac. org/ reports/ provisional/ abstract04/ connelly_310804. html) (Draft March 2004 ed.). IUPAC. 2004. p.16. .

Trans-spanning ligand
Trans-spanning ligands are bidentate ligands that can span opposite sites of a complex with square-planar geometry. A wide variety of ligands that chelate in the cis fashion already exist, but very few can link opposite vertices on a coordination polyhedron. Early attempts to generate trans-spanning bidentate ligands relied on polymethylene chains to link the donor functionalities, but such ligands often lead to coordination polymers.

History
A diphosphane linked with pentamethylene was claimed to span across a square planar complex. This early attempt was followed by ligands with more rigid backbones. "TRANSPHOS" was the first trans-spanning diphosphane ligand that usually coordinates to palladium(II) and platinum(II) in a trans manner. TRANSPHOS features benzo[c]phenanthrene substituted by diphenylphosphinomethyl (Ph2PCH2) groups at the 1 and 11 positions.[1][2] The polycyclic framework suffers sterically clashing hydrogen centers.

Trans-spanning ligand

24

Xantphos, SPANphos, TRANSDIP and related ligands


Xantphos is a trans-spanning ligand, without the steric problems associated with TRANSPHOS. SPANphos is comparable to XANTPHOS but more reliably trans-spanning. TRANSDIP, based on a -cyclodextrin, is the first ligand to give exclusively trans-spanned complexes, even with d8 metal ion halides.[3]

References
[1] Destefano, N. J.; Johnson, D. K.; Lane, R. M.; Venanzi, L. M. (1976). "Transition-Metal Complexes with Bidentate Ligands Spanning Trans-Positions. I. The Synthesis of 2,11-Bis(Diphenylphosphinomethyl)Benzo[C]-Phenanthrene, A Ligand Promoting the Formation of Square Planar Complexes". Helvetica Chimica Acta 59 (8): 26742682. doi:10.1002/hlca.19760590806. [2] Mochida, J. A.; Mattern, J. C.; Bailar, Jr., J. (1975). "Stereochemistry of Complex Inorganic Compounds. XXXV. A Complex Containing a Ligand That Spans Trans Positions". J. Am. Chem. Soc. 97 (11): 30213026. doi:10.1021/ja00844a017. [3] Poorters, L.; Armspach, D.; Dominique, M.; Toupet, L.; Choua, S.; Turek, P. (2007). "Synthesis and Properties of Transdip, A Rigid Chelator Built Upon a Cyclodextrin Cavity: is Transdip an Authentic Trans - Spanning Ligand?". Chemistry: A European Journal 13 (34): 94489461. doi:10.1002/chem.200700831. PMID17943701.

Linkage isomerism
Linkage isomerism is the existence of co-ordination compounds that have the same composition differing with the connectivity of the metal to a ligand. Typical ligands that give rise to linkage isomers are: thiocyanate, SCNselenocyanate, SeCNnitrite, NO2sulfite, SO32-

Examples of linkage isomers are violet-colored [(NH3)5Co-SCN]2+ and orange-colored [(NH3)5Co-NCS]2+. The isomerization of the S-bonded isomer to the N-bonded isomer occurs intramolecularly.[1] In the complex, dichlorotetrakis(dimethyl sulfoxide)ruthenium(II), linkage isomerism of dimethyl sulfoxide ligands can be observed in the NMR spectrum due to the effect of S vs. O bonding on the methyl groups of DMSO. The proper notation for linkage isomerism is the kappa notation where the atom directly bonding to the metal is proceeded by the lowercase Greek letter kappa; . For example, NO2- is represented as nitrito--N and nitrito--O, replacing the old system of trivial names such as nitro and nitroso.[2]

History
The first reported example of linkage isomerism had the formula [Co(NH3)5(NO2)]Cl2. The cationic cobalt complex exists in two separable linkage isomers. In the yellow-coloured isomer, the nitro ligand is bound through nitrogen. In the red linkage isomer, the nitrito is bound through one oxygen atom. The O-bonded isomer is often written as [Co(NH3)5(ONO)]2+. Although the existence of the isomers had been known since the late 1800s, only in 1907 was the structural difference explained.[3] It was later shown that the red isomer converted to the yellow isomer upon UV-irradiation. In this particular example, the formation of the nitro isomer (Co-NO2) from the nitrito isomer (Co-ONO) occurs through the rearrangement of the molecular structure. Thus, no bonds are broken during isomerization.

Linkage isomerism

25

Structures of the two linkage isomers of [Co(NH3)5(NO2)]2+.

References
[1] Buckingham, D. A.; Creaser, I. I.; Sargeson, A. M. (1970). "Mechanism of Base Hydrolysis for CoIII(NH3)5X2+ Ions. Hydrolysis and Rearrangement for the Sulfur-Bonded Co(NH3)5SCN2+ Ion". Inorg. Chem. 9 (3): 655661. doi:10.1021/ic50085a044.< [2] IUPAC. Compendium of Chemical Terminology, 2nd ed. (the "Gold Book"). Compiled by A. D. McNaught and A. Wilkinson. Blackwell Scientific Publications, Oxford (1997). XML on-line corrected version: http:/ / goldbook. iupac. org (2006-) created by M. Nic, J. Jirat, B. Kosata; updates compiled by A. Jenkins. ISBN 0-9678550-9-8. [3] Werner, A. (1907). "ber strukturisomere Salze der Rhodanwasserstoffsure und der salpetrigen Sure". Ber. 40 (1): 765788. doi:10.1002/cber.190704001117.

Bridging ligand
A bridging ligand is a ligand that connects two or more atoms, usually metal ions.[1] The ligand may be atomic or polyatomic. Virtually all complex organic compounds can serve as bridging ligands, so the term is usually restricted to small ligands such as pseudohalides or to ligands that are specifically designed to link two metals. In naming a complex wherein a single atom bridges two metals, the bridging ligand is preceded by the Greek character 'mu', [2], with a superscript number denoting the number of metals bound to the bridging ligand is bound. 2 is often denoted simply as .

Illustrative bridging ligands


Virtually all ligands are known to bridge, with the exception of amines and ammonia.[3] Particularly common inorganic bridging ligands include OH, O2, S2, SH, NH2 NH2 (imido) N3 (nitrido) CO Halides Hydride Cyanide
An example of a 2 bridging ligand

Bridging ligand Cyanide usually bridges via M-NC-M' linkages, unlike the other entries on this list. Many organic ligands form strong bridges between metal centers. Many common examples include derivatives of the above inorganic ligands (R = alkyl, aryl): OR, SR, NR2 NR2 (imido) P3 (phosphido) PR2 (phosphido, note the ambiguity with the preceding entry) PR2 (phosphinidino)

26

Polyfunctional ligands
Polyfunctional ligands can attach to metals in many ways and thus can bridge metals in diverse ways, including sharing of one atom or using several atoms. Examples of such polyatomic ligands are the oxoanions CO32 and the related Carboxylate, PO43, and the polyoxometallates. Several organophosphorus ligands have been developed that bridge pairs of metals, a well-known example being Ph2PCH2PPh2.

Examples
Compound Formula Description

{(Fe(III)(OH2)4)2(-OH)2}4+ In this example hydroxide plays the role of a 2 bridging ligand. Notice in the name of the compound 2 has been [4] simplified to .

(6-C6H6)2Ru2Cl2(-Cl)2

In this particular complex, two chloride ligands are terminal and two are 2 bridging. The in the beginning of the formula denotes the hapticity of the benzene ligands.

B2H6

This classic borane compound, diborane features two 2 bridging hydrides.

Bridging ligand

27

(Co(CO)3)3(3-(C-tBu))

This compound contains a 3 bridging carbyne ligand (C-tBu).

See Also
Bridging carbonyl

References
[1] Nic, M.; Jirat, J.; Kosata, B., eds. (2006). "bridging ligand" (http:/ / goldbook. iupac. org/ B00741. html). IUPAC Compendium of Chemical Terminology (Online ed.). doi:10.1351/goldbook.{{{file}}}. ISBN0-9678550-9-8. . [2] Nic, M.; Jirat, J.; Kosata, B., eds. (2006). "- (mu)" (http:/ / goldbook. iupac. org/ M03659. html). IUPAC Compendium of Chemical Terminology (Online ed.). doi:10.1351/goldbook.{{{file}}}. ISBN0-9678550-9-8. . [3] Werner, H. (2004). "The Way into the Bridge: A New Bonding Mode of Tertiary Phosphanes, Arsanes, and Stibanes". Angew. Chem. Int. Ed. 43 (8): 938954. doi:10.1002/anie.200300627. PMID14966876. [4] Classifications of ligands. http:/ / chimge. unil. ch/ En/ complexes/ 1cpx7. htm

Metalligand multiple bond


In Chemistry, a metalligand multiple bond describes the interaction of certain ligands with a metal with a bond order greater than one.[1] Coordination complexes featuring multiply bonded ligands are of both scholarly and practical interest. Transition metal carbene complexes catalyze the olefin metathesis reaction. Metal oxo intermediates are pervasive in oxidation catalysis. oxygen evolving complex. As a cautionary note, the classification of a metal ligand bond as being "multiple" bond order is ambiguous and even arbitrary because bond order is a formalism. Furthermore, the usage of multiple bonding is not uniform. Symmetry arguments suggest that most ligands engage metals via multiple bonds. The term 'metal ligand multiple bond" is often reserved for ligands of the type CRn and NRn (n = 0, 1, 2) and ORn (n = 0, 1) where R is H or an organic substituent, or pseudohalide. Historically, CO and NO+ are not included in this classification, nor are halides.

Most common classes of complexes showing metal ligand multiple bonds

Metalligand multiple bond

28

Pi-donor ligands
In coordination chemistry, a pi-donor ligand is a kind of ligand endowed with filled non-bonding orbitals that overlap with metal-based orbitals. Their interaction is complementary to the behavior of pi-acceptor ligands. The existence of terminal oxo ligands for the early transition metals is one consequence of this kind of bonding. Classic pi-donor ligands are oxide (O2-), nitride (N3-), imide (RN2-), alkoxide (RO-), amide (R2N-), and fluoride. For late transition metals, strong pi-donors form anti-bonding interactions with the filled d-levels, with consequences for spin state, redox potentials, and ligand exchange rates. Pi-donor ligands are low in the spectrochemical series.[1]

Multiple bond stabilization


Metals bound to so-called triply bonded carbyne, imide, nitride (nitrido), and oxide (oxo) ligands are generally assigned to high oxidation states with low d electron counts. The high oxidation state stabilizes the highly reduced ligands. The low d electron count allow for many bonds between ligands and the metal center. A d0 metal center can accommodate up to 9 bonds without violating the 18 electron rule, whereas a d6 species can only accommodate 6 bonds.

Reactivity explained through ligand hybridization


A ligand described in ionic terms can bond to a metal through however many lone pairs it has available. For example many alkoxides use one of their three lone pairs to make a single bond to a metal center. In this situation the oxygen is sp3 hybridized according to valence bond theory. Increasing the bond order to two by involving another lone pair changes the hybridization at the oxygen to an sp2 center with an expected expansion in the M-O-R bond angle and contraction in the M-O bond length. If all three lone pairs are included for a bond order of three than the M-O bond distance contracts further and since the oxygen is a sp center the M-O-R bond angle is 180 or linear. Similarly with the imidos are commonly referred to as either bent (sp2) or linear (sp). Even the oxo can be sp2 or sp hybridized. The triply bonded oxo, similar to carbon monoxide, is partially positive at the oxygen atom and unreactive towards bronsted acids at the oxygen atom. When such a complex is reduced, the triple bond can be converted to a double bond at which point the oxygen no longer bears a partial positive charge and is reactive towards acid.

Conventions
Bonding representations
Imido ligands, also known as imides or nitrenes, most commonly form "linear six electron bonds" with metal centers. Bent imidos are a rarity limited by complexes electron count, orbital bonding availability, or some similar phenomenon. It is common to draw only two lines of bonding for all imidos, including the most common linear imidos with a six electron bonding interaction to the metal center. Similarly amido complexes are usually drawn with a single line even though most amido bonds involve four electrons. Alkoxides are generally drawn with a single bond although both two and four electron bonds are common. Oxo can be drawn with two lines regardless of whether four electrons or six are involved in the bond, although it is not uncommon to see six electron oxo bonds represented with three lines.

Representing oxidation states


There are two motifs to indicate a metal oxidation state based around the actual charge separation of the metal center. Oxidation states up to +3 are believed to be an accurate representation of the charge separation experienced by the metal center. For oxidation states of +4 and larger, the oxidation state becomes more of a formalism with much of the positive charge distributed between the ligands. This distinction can be expressed by using a Roman numeral for the lower oxidation states in the upper right of the metal atomic symbol and an Arabic number with a

Metalligand multiple bond plus sign for the higher oxidation states (see the example below). This formalism is not rigorously followed and the use of Roman numerals to represent higher oxidation states is common. [MIIILn]3+ vs. [O=M5+Ln]3+

29

References
[1] "MetalLigand Multiple Bonds: The Chemistry of Transition Metal Complexes Containing Oxo, Nitrido, Imido, Alkylidene, or Alkylidyne Ligands" W. A. Nugent and J. M. Mayer; Wiley-Interscience, New York, 1988.

Further reading (specialized literature)


Heidt, L.J.; Koster, G.F.; Johnson, A.H. "Experimental and Crystal Field Study of the Absorption Spectrum at 2000 to 8000 A of to Manganous Perchlorate in Aqueous Perchloric Acid" J. Am. Chem. Soc. 1959, 80, 64716477. Rohde,J; In,J.; Lim, M.H.; Brennessel, W.W.; Bukowski, M.R.; Stubna, A.; Muonck, E.; Nam, W.; Que L. "Crystallographic and Spectroscopic Characterization of a Nonheme Fe(IV)O Complex" Science VOL 299 10371039. Decker, A.; Rohde,J.; Que, L.; Solomon, E.I. "Spectroscopic and Quantum Chemical Characterization of the Electronic Structure and Bonding in a Non-Heme FeIVO Complex" J. Am. Chem. Soc. 2004, 126, 53785379. Aliaga-Alcalde, N.; George, S.D.; Mienert, B.; Bill, E.; Wieghardt, K.; Neese, F. "The Geometric and Electronic Structure of [(cyclam-acetato)Fe(N)]+: A Genuine Iron(V) Species with a Ground-State Spin S=1/2" Angew. Chem. Int. Ed. 2005, 44, 29082912.

Non-innocent ligand
In chemistry, a (redox) non-innocent ligand is a ligand in a metal complex where the oxidation state is unclear. Typically, complexes containing non-innocent ligands are redox active at mild potentials. The concept assumes that redox reactions in metal complexes are either metal or ligand localized, which is a simplification, albeit a useful one.

Redox Reactions of Complexes Containing Innocent vs. Non-Innocent Ligands


Conventionally, redox reactions of coordination complexes are assumed to be metal-centered. The reduction of MnO4- to MnO42- is described by the change in oxidation state of manganese from 7+ to 6+. The oxide ligands do not change in oxidation state, remaining 2- (a more careful examination of the electronic structure of the redox partners reveals however that the oxide ligands are affected by the redox change). Oxide is an innocent ligand. Another example of conventional metal-centered redox couple is [Co(NH3)6]3+/[Co(NH3)6]2+. Ammonia is innocent in this transformation.

A clear example of redox non-innocent behavior of ligands is observed for [Ni(S2C2Ph2)2]z, which exists in three oxidation states: z = 2-, 1-, and 0. If the ligands are always considered to be dianionic (as is done in formal oxidation state counting), then z = 0 requires that that nickel has a formal oxidation state of +IV. The formal oxidation state of the central nickel atom therefore ranges from +II to +IV in the above transformations (see Figure). However, the formal oxidation state is different from the real (spectroscopic) oxidation state based on the (spectroscopic) metal

Non-innocent ligand d-electron configuration. The stilbene-1,2-dithiolate behaves as a redox non-innocent ligand, and the oxidation processes actually take place at the ligands rather than the metal. This leads to the formation of ligand radical complexes. The charge-neutral complex (z =0) is therefore best described as a Ni2+ derivative of S2C2Ph2-. The diamagnetism of this complex arises from anti-ferromagnetic coupling between the unpaired electrons of the two ligand radicals. The complex Cr(2,2'-bipyridine)3 is a derivative of Cr(III) bound to three radical anions of 2,2'bipyridine, which is in this case also behaving as a redox non-innocent ligand. On the other hand, one-electron oxidation of [Ru(2,2'-bipyridine)3]2+ is localized on Ru and the bipyridine is behaving as a normal, innocent ligand in this case.

30

History
C.K. Jrgenson (Cologny-Geneva) described ligands as "innocent" and "suspect": "Ligands are innocent when they allow oxidation states of the central atoms to be defined. The simplest case of a suspect ligand is NO..."[1] Redox non-innocent ligands have been intensively investigated spectroscopically by the groups of K. Wieghardt (MPI Mlheim a/d Ruhr) and W. Kaim (Stuttgart) over the past years. Quite recently it became obvious that redox non-innocent ligands are not just a spectroscopic curiosity, as the radical reactivity of redox non-innocent ligands was demonstrated to play a crucial role in the mechanism of bio-catalytic processes mediated by several metallo-enzymes (e.g. Gallactose Oxidase, Cytochrome P450, methane mono-oxygenase). More recently, some synthetic research groups have started to systematically investigate the (catalytic) reactivity of transition metal complexes with redox non-innocent ligands in organometallic chemistry.

Typical Ligands that often behave as Redox Non-Innocent Ligands


O2 and NO.[2] Ligands with extended pi-delocalization such as porphyrins and phthalocyanines, ligands with the generalised formulas [D-CR=CR-D]2- or D=CR-CR=D (D = O, S, NR and R, R' = alkyl or aryl), and similar related systems are often non-innocent. For example: dioxalenes, such as catecholates.[3] dithiolenes, such as 1,2-maleonitriledithiolate diimines such as derivatives of 1,2-diaminobenzene, -diimines, and dimethylglyoxime. pyridine-2,6-diimine ligands (relevant in polymerisation and hydrogenation catalysis).

The pyridine-2,6-diimine ligand can be easily reduced by one or two electrons.[4][5][6]

Non-innocent ligand

31

Redox Non-Innocent Ligands in Organometallic Chemistry and Catalysis


In paramagnetic organometallic complexes of Rh and Ir (metallo-radicals),[7] ethene ligands, amido ligands, and (reactive) carbene ligands are sometimes also behaving as 'redox non-innocent' ligands:

Solvent coordination to some metallo-radical IrII(ethene) species transfers the spin-density from the metal to the redox non-innocent ethene ligand, after which direct radical coupling reactions with the olefinic ligand radical become possible.[8][9][10] Oxidation of certain RhI-amido and IrI-amido complexes does not lead to the expected MII-amido species. Instead the unexpected MI-aminyl radical complexes are formed.[11] Carbene formation from diazo compounds at metallo-radical IrII species unexpectedly leads to formation of 'carbene radicals'. This is a result of the redox non-innocent character of Fischer-type carbenes, where one-electron reduction of the carbene ligand by IrII leads to formation of carbon centered 'carbene radicals' coordinated to IrIII. These 'carbene radicals' reveal interesting radical-type reactivities.[12]

Non-innocent ligand

32

Redox Non-Innocent Ligands in Biology


Metalloenzymes often feature non-innocent ligands. A common non-innocent ligand is found in metalloporphyrins. In the enzyme cytochrome P450, the porphyrin ligand sustains oxidation during the catalytic cycle. In other heme proteins, such as myoglobin, ligand-centered redox does not occur and the porphyrin is innocent.

Galactose Oxidase (GOase) provides a seminal example for the involvement of reactive non-innocent ligands in bio-catalytic turnover.[13][14] GOase converts chemo-selectively primary alcohols with O2 into aldehydes and H2O2, with impressive turnover frequencies. The active site of the enzyme GOase contains a tyrosinyl radical which is coordinated to a CuII ion. In the key steps of the catalytic cycle, a cooperative Brnsted-basic ligand-site deprotonates the alcohol, and subsequently the oxygen atom of the tyrosinyl radical abstracts a hydrogen atom from the alpha-CH functionality of the coordinated alcoholate substrate. Thus, the tyrosinyl radical is a reactive fragment in the catalytic cycle which cooperates with the Cu site. This is essential for the function of the enzyme, because the Cu-ion is only capable of one-electron transformations. It is the interplay of the 1e reactivity of the ligand radical and the 1e reactivity of the metal which makes the overall process possible. The radical abstraction nature of the process makes the process extremely fast. Anti-ferromagnetic coupling between the unpaired spins of the tyrosine radical ligand and the d9 CuII ion (open-shell singlet ground state) explains the observed diamagnetic nature of the resting state of the enzyme, as was confirmed by synthetic model studies.[15] The oxygen molecule in oxyhemoglobin (or oxymyoglobin) would appear to satisfy the definition of a non-innocent ligand. Deoxyhemoglobin is ferrous and pentacoordinate, the (innocent) ligands being four N of the porphyrin and Ne of the proximal histidine. An O2 molecule binds the sixth coordination position. There is evidence from a number of lines that partial electron transfer from the Fe to O2 occurs, so that the complex is better described as superoxide anion bound to ferric heme [16], although spin coupling makes the complex diamagnetic. The change in oxidation and spin state of the Fe results in a change of bond length to the five innocent ligands which results in the heme switching from a "domed" to a planar conformation, which in turn drives conformational changes in the protein responsible for the cooperativity of O2 binding. This cooperativity is essential for efficient oxygen transport, so in a way we all owe our lives to the suspect nature of the O2 ligand!

Non-innocent ligand

33

References
[1] Jrgensen, Chr. K. (1966). "Differences between the four halide ligands, and discussion remarks on trigonal-bipyramidal complexes, on oxidation states, and on diagonal elements of one-electron energy". Coordination Chemistry Reviews 1 (1-2): 164178. doi:10.1016/S0010-8545(00)80170-8. [2] Kaim, W.; Schwederski, B. (2010). "Non-innocent ligands in bioinorganic chemistryAn overview". Coordination Chemistry Reviews. 254 (13-14) (13-14): 15801588. doi:10.1016/j.ccr.2010.01.009. [3] Piero Zanello, P.; Corsini, M. (2006). "Homoleptic, mononuclear transition metal complexes of 1,2-dioxolenes: Updating their electrochemical-to-structural (X-ray) properties". Coordination Chemistry Reviews 250 (15-16): 20002022. doi:10.1016/j.ccr.2005.12.017. [4] de Bruin, B.; Bill, E.; Bothe, E.; Weyhermller, T.; Wieghardt, K. (2000). "Molecular and Electronic Structures of Bis(pyridine-2,6-diimine)metal Complexes [ML2](PF6)n(n = 0, 1, 2, 3; M = Mn, Fe, Co, Ni, Cu, Zn)". Inorganic Chemistry 39 (13): 29362947. doi:10.1021/ic000113j. [5] Budzelaar, P.H.M.; de Bruin, B.; Gal, A.W.; Wieghardt, K.; van Lenthe, J.H. (2001). "Metal-to-Ligand Electron Transfer in Diiminopyridine Complexes of MnZn. A Theoretical Study". Inorganic Chemistry 40 (18): 46494655. doi:10.1021/ic001457c. [6] Chirik, P.J.; Wieghardt, K. (2010). "Radical Ligands Confer Nobility on Base-Metal Catalysts". Science. 327 (5967) (5967): 794795. doi:10.1126/science.1183281. PMID20150476. [7] de Bruin, B.; Hetterscheid, D.G.H.; Koekkoek, A.J.J.; Grtzmacher, H. (2007). 5. In Karlin, Kenneth D.. "The Organometallic Chemistry of Rh-, Ir-, Pd-, and Pt-Based Radicals: Higher Valent Species". Progress in Inorganic Chemistry 55: 247354. doi:10.1002/9780470144428. [8] Hetterscheid, D.G.H.; Kaiser, J.; Reijerse, E.; Peters, T.P.J.; Thewissen, S.; Blok, A.N.J.; Smits, J.M.M.; de Gelder, R.; de Bruin, B. (2005). "IrII(ethene): Metal or Carbon Radical?". Journal of the American Chemical Society 127 (6): 18951905. doi:10.1021/ja0439470. PMID15701024. [9] Hetterscheid, D.G.H.; Bens, M.; de Bruin, B. (2005). "IrII(ethene): Metal or Carbon Radical? Part II: Oxygenation via iridium or direct oxygenation at ethene?". Dalton Transactions (5): 979984. doi:10.1039/b417766e. PMID15726153. [10] de Bruin, B.; Hetterscheid, D.G.H. (2007). "Paramagnetic (Alkene)Rh and (Alkene)Ir Complexes: Metal or Ligand Radicals?". European Journal of Inorganic Chemistry 2 (2): 211230. doi:10.1002/ejic.200600923. [11] Bttner, T.; Geier, J.; Frison, G.; Harmer, J.; Calle, C.; Schweiger, A.; Schnberg, H.; Grtzmacher, H. (2005). "A Stable Aminyl Radical Metal Complex". Science. 307 (5707) (5707): 235238. doi:10.1126/science.1106070. PMID15653498. [12] Dzik, W.I.; Reek, J.N.H.; de Bruin, B. (2008). "Selective C-C Coupling of Ir-Ethene and Ir-Carbenoid Radicals". Chemistry: A European Journal 14 (25): 75947599. doi:10.1002/chem.200800262. PMID18523935. [13] Whittaker, M.M.; Whittaker, J.W. (1993). "Ligand interactions with galactose oxidase: mechanistic insights". Biophysical Journal. 64 (3): 762772. [14] Wang, Y.; DuBois, J. L.; Hedman, B.; Hodgson, K. O.; Stack, T. D. P. (1998). "Catalytic Galactose Oxidase Models: Biomimetic Cu(II)-Phenoxyl-Radical Reactivity". Science. 279 (5350) (5350): 537540. doi:10.1126/science.279.5350.537. [15] Mller, J.; Weyhermller, T. Bill, E.; Hildebrandt, P.; Ould-Moussa, L.; Glaser, T.; Wieghardt, K. (1998). "Why Does the Active Form of Galactose Oxidase Possess a Diamagnetic Ground State?". Angewandte Chemie International Edition. 37 (5) (5): 616619. doi:10.1002/(SICI)1521-3773(19980316)37:5<616::AID-ANIE616>3.0.CO;2-4. [16] Hui Chen, Masao Ikeda-Saito and Sason Shaik (2008). "Nature of the FeO2 Bonding in Oxy-Myoglobin: Effect of the Protein". J. Am. Chem. Soc. (130): 1477814790. doi:10.1021/ja805434m.

Chiral ligand

34

Chiral ligand
In chemistry a chiral ligand is a specially adapted ligand used for asymmetric synthesis. This ligand is an enantiopure organic compound which combines with a metal center by chelation to form an asymmetric catalyst. This catalyst engages in a chemical reaction and transfers its chirality to the reaction product which as a result also becomes chiral. In an ideal reaction one equivalent of catalyst can turn over many more equivalents of reactant which enables the synthesis of a large amount of a chiral compound from achiral precursors with the aid of a very small (often expensive) chiral ligand.

First discovery
The first such ligand, the diphosphine DiPAMP was developed in 1968 by William S. Knowles of Monsanto Company, who won the 2001 Nobel Prize in Chemistry,[1] and ultimately used in the industrial production of L-DOPA.

Privileged ligands
Many thousands of chiral ligands have been prepared and tested since then but only several compound classes have been found to have a general scope. These ligands are therefore called privileged ligands.[2][3] Important members depicted below are BINOL, BINAP, TADDOL, DIOP, BOX and DuPhos (a phosphine ligand), all available as enantiomeric pairs.

Other members are Salen, cinchona alkaloids and phosphoramidites. Many of these ligands possess C2 symmetry which limits the number of possible reaction pathways and thereby increasing enantioselectivity.

Chiral ligand

35

Chiral fence
Chiral ligands work asymmetric induction somewhere along the reaction coordinate. The image depicted on the right gives a general idea how a chiral ligand may induce an enantioselective reaction. The ligand (in green) has C2 symmetry with its nitrogen, oxygen or phosphorus atoms hugging a central metal atom (in red). In this particular ligand the right side is sticking out and its left side points away. The substrate in this reduction is acetophenone and the reagent (in blue) a hydride ion. In absence of the metal and the ligand the re face approach of the hydride ion gives the (S)-enantiomer and the si face approach the (R)-enantiomer in equal amounts (a racemic mixture like expected). The ligand/metal presence changes all that. The carbonyl group will coordinate with the metal and due to the steric bulk of the phenyl group it will only be able to do so with its si face exposed to the hydride ion with in the ideal situation exclusive formation of the (R) enantiomer. The re face will simple hit the chiral fence.[4] Note that when the ligand is replaced by its mirror image the other enantiomer will form and that a racemic mixture of ligand will once again yield a racemic product. Also note that if the steric bulk of both carbonyl substituents is very similar the strategy will fail.

Chiral ligand

36

Chiral counterions
In a novel concept, so-called chiral ions team up with traditional cationic catalysts in asymmetric synthesis as demonstrated in this allene hydroxyalkoxylation in which the active catalyst is a salt of gold(I) and a phosphate of a chiral binaphthol:[5][6]

References
[1] Nobel prize 2001 www.nobelprize.org Link (http:/ / nobelprize. org/ nobel_prizes/ chemistry/ laureates/ 2001/ public. html) [2] Design of chiral ligands for asymmetric catalysis: From C2-symmetric P,P- and N,N-ligands to sterically and electronically nonsymmetrical P,N-ligands Andreas Pfaltz and William J. Drury III PNAS, April 20, 2004 vol. 101 no. 16 5723-5726 doi:10.1073/pnas.0307152101 [3] Privileged Chiral Catalysts Tehshik P. Yoon, Eric N. Jacobsen Science 14 March 2003: Vol. 299. no. 5613, pp. 1691 - 1693 doi:10.1126/science.1083622 PMID 12637734 [4] Chiral and C2-symmetrical bis(oxazolinylpyridine)rhodium(III) complexes: effective catalysts for asymmetric hydrosilylation of ketones Hisao Nishiyama, Hisao Sakaguchi, Takashi Nakamura, Mihoko Horihata, Manabu Kondo, and Kenji Itoh Organometallics; 1989; 8(3) pp 846 - 848; doi: 10.1021/om00105a047 [5] A Powerful Chiral Counterion Strategy for Asymmetric Transition Metal Catalysis Gregory L. Hamilton, Eun Joo Kang, Miriam Mba, F. Dean Toste. Science 317, 496 (2007) doi: 10.1126/science.1145229 [6] Starting catalyst: 1,2-bis(diphenylphosphino)ethane (dppm) gold(I) chloride complex

Ligand dependent pathway

37

Ligand dependent pathway


There are two types of pathway for substitution of ligands in a complex. The ligand dependent pathway is the one whereby the chemical properties of the ligand affect the rate of substitution. Alternatively, there is the ligand independent pathway, which is where the ligand does not have an effect. This is of vital importance in the world of inorganic chemistry and complex ions.

Ligand field theory


Ligand field theory (LFT) describes the bonding, orbital arrangement, and other characteristics of coordination complexes.[1] It represents an application of molecular orbital theory to transition metal complexes. A transition metal ion has nine valence atomic orbitals, five (n)d, one (n+1)s, and three (n+1)p orbitals. These orbitals are of appropriate energy to form bonding interaction 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 to the metal.[2]

-Bonding
The molecular orbitals created by coordination can be seen as resulting from the donation of two electrons by each of six -donor ligands to the d-orbitals on the metal. In octahedral complexes, ligands approach along the x-, y- and z-axes, so their -symmetry orbitals form bonding and anti-bonding combinations with the dz2 and dx2y2 orbitals. The dxy, dxz and dyz orbitals remain non-bonding orbitals. Some weak bonding (and anti-bonding) interactions with the s and p orbitals of the metal also occur, to make a total of 6 bonding (and 6 anti-bonding) molecular orbitals.

Ligand-Field scheme summarizing -bonding in the octahedral complex [Ti(H2O)6]3+.

Ligand field theory In molecular symmetry terms, the six lone pair orbitals from the ligands (one from each ligand) form six symmetry adapted linear combinations (SALCs) of orbitals, also sometimes called ligand group orbitals (LGOs). The irreducible representations that these span are a1g, t1u and eg. The metal also has six valence orbitals that span these irreducible representations - the s orbital is labeled a1g, a set of three p-orbitals is labeled t1u, and the dz2 and dx2y2 orbitals are labeled eg. The six -bonding molecular orbitals result from the combinations of ligand SALC's with metal orbitals of the same symmetry.

38

-bonding
bonding in octahedral complexes occurs in two ways: via any ligand p-orbitals that are not being used in bonding, and via any or * molecular orbitals present on the ligand. The p-orbitals of the metal are used for bonding (and are the wrong symmetry to overlap with the ligand p or or * orbitals anyway), so the interactions take place with the appropriate metal d-orbitals, i.e. dxy, dxz and dyz. These are the orbitals that are non-bonding when only bonding takes place. One important bonding in coordination complexes is metal-to-ligand bonding, also called backbonding. It occurs when the LUMOs of the ligand are anti-bonding * orbitals. These orbitals are close in energy to the dxy, dxz and dyz orbitals, with which they combine to form bonding orbitals (i.e. orbitals of lower energy than the aforementioned set of d-orbitals). The corresponding anti-bonding orbitals are higher in energy than the anti-bonding orbitals from bonding so, after the new bonding orbitals are filled with electrons from the metal d-orbitals, O has increased and the bond between the ligand and the metal strengthens. The ligands end up with electrons in their * molecular orbital, so the corresponding bond within the ligand weakens. The other form of coordination bonding is ligand-to-metal bonding. This situation arises when the -symmetry p or orbitals on the ligands are filled. They combine with the dxy, dxz and dyz orbitals on the metal and donate electrons to the resulting -symmetry bonding orbital between them and the metal. The metal-ligand bond is somewhat strengthened by this interaction, but the complementary anti-bonding molecular orbital from ligand-to-metal bonding is not higher in energy than the anti-bonding molecular orbital from the bonding. It is filled with electrons from the metal d-orbitals, however, becoming the HOMO of the complex. For that reason, O decreases when ligand-to-metal bonding occurs. The greater stabilisation that results from metal-to-ligand bonding is caused by the donation of negative charge away from the metal ion, towards the ligands. This allows the metal to accept the bonds more easily. The combination of ligand-to-metal -bonding and metal-to-ligand -bonding is a synergic effect, as each enhances the other. As each of the six ligands has two orbitals of -symmetry, there are twelve in total. The symmetry adapted linear combinations of these fall into four triply degenerate irreducible representations, one of which is of t2g symmetry. The dxy, dxz and dyz orbitals on the metal also have this symmetry, and so the -bonds formed between a central metal and six ligands also have it (as these -bonds are just formed by the overlap of two sets of orbitals with t2g symmetry.)

High and low spin and the spectrochemical series


The six bonding molecular orbitals that are formed are "filled" with the electrons from the ligands, and electrons from the d-orbitals of the metal ion occupy the non-bonding and, in some cases, anti-bonding MO's. The energy difference between the latter two types of MO's is called O (O stands for octahedral) and is determined by the nature of the -interaction between the ligand orbitals with the d-orbitals on the central atom. As described above, -donor ligands lead to a small O and are called weak- or low-field ligands, whereas -acceptor ligands lead to a large value of O and are called strong- or high-field ligands. Ligands that are neither -donor nor -acceptor give a value of O somewhere in-between.

Ligand field theory The size of O determines the electronic structure of the d4 - d7 ions. In complexes of metals with these d-electron configurations, the non-bonding and anti-bonding molecular orbitals can be filled in two ways: one in which as many electrons as possible are put in the non-bonding orbitals before filling the anti-bonding orbitals, and one in which as many unpaired electrons as possible are put in. The former case is called low-spin, while the latter is called high-spin. A small O can be overcome by the energetic gain from not pairing the electrons, leading to high-spin. When O is large, however, the spin-pairing energy becomes negligible by comparison and a low-spin state arises. The spectrochemical series is an empirically-derived list of ligands ordered by the size of the splitting that they produce. It can be seen that the low-field ligands are all -donors (such as I-), the high field ligands are -acceptors (such as CN- and CO), and ligands such as H2O and NH3, which are neither, are in the middle. I < Br < S2 < SCN < Cl < NO3 < N3 < F < OH < C2O42 < H2O < NCS < CH3CN < py (pyridine) < NH3 < en (ethylenediamine) < bipy (2,2'-bipyridine) < phen (1,10-phenanthroline) < NO2 < PPh3 < CN < CO

39

History
Ligand field theory was developed during the 1930s and 1940s as an alternative to crystal field theory (CFT). CFT describes certain properties of coordination complexes but is based on a model that emphasizes electrostatic interactions between ligand electrons with the d-electrons on the metal. CFT does not describe bonding. Ligand Field Theory, in a sense, combined CFT and the then-emerging molecular orbital theory.

References
[1] Schlfer, H. L.; Gliemann, G. "Basic Principles of Ligand Field Theory" Wiley Interscience: New York; 1969 [2] G. L. Miessler and D. A. Tarr Inorganic Chemistry 3rd Ed, Pearson/Prentice Hall publisher, ISBN 0-13-035471-6.

Article Sources and Contributors

40

Article Sources and Contributors


Ligand Source: http://en.wikipedia.org/w/index.php?oldid=502666113 Contributors: 194.200.130.xxx, 1exec1, 4lex, Ace Frahm, Achemgeekf, Ahoerstemeier, Akhilchirravuri, Alexnullnullsieben, AlphaEta, Altenmann, Amangill, Amateria1121, AnnaFrance, Apple2, Ausinha, Ave matthew, Axiosaurus, Basicdesign, Beetstra, Benbest, Benjah-bmm27, Bhellis, Biscuittin, Bissinger, Brandon, Brichcja, Bryan Derksen, CMSherwood, Cadmium, Centrx, Ceyockey, Chem8240bv, ChemMater, Christian75, Cmrufo, Conrad.Irwin, Conversion script, Crystal whacker, Cstmoore, Dcirovic, Dilbert08, Dmb000006, Dormroomchemist, Drphilharmonic, Dualus, Dwmyers, Erajda, Fnielsen, Gadfium, Gaius Cornelius, Garethhegarty, Gentgeen, Giftlite, Gilliam, Gioto, Habj, Hugo-cs, Itub, Jrissman, Jrtayloriv, Junior Brian, Jynto, K. Hiippari, Kenster85hero, Ktsquare, Lfh, Lifeformnoho, Lkinkade, Lus Felipe Braga, Magwich77, Manderson198, Martarius, Mattbr, Mayooranathan, Mets501, Michael Hardy, Mj455972007, Mrnatural, Nihiltres, Nina Gerlach, Nlu, Nono64, OMCV, Organometallics, Patrick Lutz, Pentalis, Petrb, Pgholder, Pixeltoo, Postglock, Pwdent, Quadalpha, Quadro, R'n'B, Red, Rej8, Rifleman 82, Sadads, Serein (renamed because of SUL), SilverShatter, SilveryWolf, Slightsmile, Smokefoot, Sodium, SrKAawa, Srnec, Stewartadcock, T.c.w7468, T.vanschaik, TDogg310, TheAMmollusc, Thegeneralguy, Thiseye, ThomasYun, Tlabshier, Tmangray, V8rik, Valenciano, Vinci cool, Walkerma, Waltpohl, Wangi, Wayne Slam, Whoop whoop pull up, Wickey-nl, Woohookitty, Yaluen, YitzHandel, Zenohockey, Zoom321, , , 152 anonymous edits Crystal field theory Source: http://en.wikipedia.org/w/index.php?oldid=505556984 Contributors: 1727quimico, 17Drew, Alansohn, Anton Khorev, Aponar Kestrel, ArgentiumOutlaw, Beetstra, Benjah-bmm27, Bit Lordy, Blethering Scot, Brichcja, CarbonX, Carsrac, Chemicalengineer03, Claviere, Dirac66, DragonflySixtyseven, Freestyle-69, Gaius Cornelius, Gmanley, Hoffmeier, J991, Jaganath, Jaraalbe, Jh51681, Kazkaskazkasako, Kingpin13, LinDrug, Manuelkuhs, Materialscientist, Michael Hardy, Moorfrogger, MrBland, OMCV, Okedem, Out of Phase User, PSU PHYS514 F06, Passw0rd, Petergans, Pettythug, PleaseStand, Prateek khanna, Puppy8800, Quadell, Quantockgoblin, RaseaC, Red Thrush, Reedy, Rifleman 82, SWpens11, Sfan00 IMG, Silent Nemesis2710, Smokefoot, Snehashistamlukwb, T.vanschaik, Tabatharose, V8rik, Walkerma, Welsh, Xiglofre, YanA, Zaiken, 78 anonymous edits Denticity Source: http://en.wikipedia.org/w/index.php?oldid=503385743 Contributors: Anypodetos, Axiosaurus, CMSherwood, CommonsDelinker, Crystal whacker, Dr. F.C. Turner, DrPhen, Imareaver, Lamro, MrBell, Rod57, Roux, Smokefoot, Wickey-nl, 6 anonymous edits Chelation Source: http://en.wikipedia.org/w/index.php?oldid=501870390 Contributors: A876, AdamRoach, Albmont, Alexei Kouprianov, Alphachimp, Anders.Warga, Andonic, Anupam, Arcadian, Archfool, AxelBoldt, Axewiki, Beetstra, Benbest, Bensaccount, Bill.albing, Bug42, Bunnyhop11, Burschik, C S, Cacycle, CaptinJohn, Cburnett, CecilWard, Centrx, Chowbok, Crystal whacker, Cuaxdon, Cutefidgety, Dcirovic, Dfranke, Dieter Simon, Dj Capricorn, Dmcarey1, Dmn, Doodle77, Drilnoth, Drsibia, Dwmyers, Elinor McCartney, Erianna, Eubulides, Filelakeshoe, Finemann, FrozenMan, Gaius Cornelius, Gene Nygaard, Gonzonoir, Gpunketesh, Grandpa Larsen, Ground Zero, HappyCamper, Herald83, Hughcharlesparker, JaGa, Jkbrown, JoJan, John Nevard, Johner, JonRichfield, Jpbrenna, Jtyndall02, Julesd, Kauczuk, Keenan Pepper, Keira Vaughn, Kelson, Kjmoran, Kristenq, Ktsquare, Kwamikagami, Langhorner, Lorenzarius, Maelli, Mboverload, Mellery, MrBell, NaOH, Narayanese, NawlinWiki, Nbarth, Neparis, Nihiltres, Ohnoitsjamie, Ojigiri, Okedem, Organometallics, Outsidelookin, Passw0rd, PaulNovitski, Persian Poet Gal, Petergans, Pgan002, Physchim62, Pinethicket, Pjacobi, Pquijal, Rajah, Ransu, Recognizance, Rifleman 82, Rjwilmsi, Sam Hocevar, Scharks, Sdel, Shaddack, Shalom Yechiel, Shunnosuke, SlamDiego, Smokefoot, Srnec, StevenDH, Sticky Parkin, StradivariusTV, TJRC, TMorris13, Talgalili, Tomas e, Tortoise0308, Trusilver, Unyoyega, Varlaam, Vuo, WLU, Westerness, Wickey-nl, Yaris678, Zymatik, , , 119 anonymous edits Hapticity Source: http://en.wikipedia.org/w/index.php?oldid=491347596 Contributors: Beetstra, Beetstra public, Brichcja, CMSherwood, Dr. Sunglasses, Edgar181, Eno-ja, Euchiasmus, Gaius Cornelius, Haeleth, HappyApple, HappyCamper, Hooperbloob, Lhynard, Nergaal, PedroDaGr8, R'n'B, RandomP, Rifleman 82, Sasuke Sarutobi, Shalom Yechiel, Shinryuu, Smokefoot, T.vanschaik, Tabletop, Tetracube, V8rik, YanA, 8 anonymous edits Trans-spanning ligand Source: http://en.wikipedia.org/w/index.php?oldid=467580083 Contributors: Christian75, Dualus, Firebat08, Itub, Lamro, OMCV, Rifleman 82, Smokefoot, Tassedethe, 1 anonymous edits Linkage isomerism Source: http://en.wikipedia.org/w/index.php?oldid=479104667 Contributors: CWenger, Charles Matthews, Christian75, Eggilicious, GeeJo, Goldenrowley, JamesBWatson, Physchim62, Puppy8800, Retropunk, Rifleman 82, Shalom Yechiel, Smokefoot, Srnec, Tiddly Tom, V8rik, 10 anonymous edits Bridging ligand Source: http://en.wikipedia.org/w/index.php?oldid=495206398 Contributors: Alan Au, Axiosaurus, Beetstra, Benjah-bmm27, CMSherwood, Chem-awb, Dslate2123, Itub, Mets501, Mikespedia, Patrick Lutz, PedroDaGr8, Physchim62, Quaxmonster, Rifleman 82, Smokefoot, Srnec, Tomsdearg92, Useight, 5 anonymous edits Metalligand multiple bond Source: http://en.wikipedia.org/w/index.php?oldid=504420935 Contributors: Chem8240bv, Dicklyon, EdChem, ErikHaugen, Hydrogen Iodide, Nergaal, OMCV, Rifleman 82, Smokefoot, SunCreator, TenPoundHammer, V8rik, Wickey-nl, 8 anonymous edits Non-innocent ligand Source: http://en.wikipedia.org/w/index.php?oldid=503504714 Contributors: Bruintje71, Eaberry, Headbomb, Nick Y., Rifleman 82, Rjwilmsi, Smokefoot, V8rik, 42 anonymous edits Chiral ligand Source: http://en.wikipedia.org/w/index.php?oldid=493770751 Contributors: AlChimini, Brossow, ChrisGualtieri, Crystal whacker, GTBacchus, Mercina87, Michael Devore, Omegakent, Rjwilmsi, Sephiroth BCR, Stever Augustus, V8rik, WereSpielChequers, 6 anonymous edits Ligand dependent pathway Source: http://en.wikipedia.org/w/index.php?oldid=471009242 Contributors: Berland, Diannaa, Malcolma, Postcard Cathy, Ronhjones, Sadads, Simondrake, Smokefoot, TexasAndroid, , 1 anonymous edits Ligand field theory Source: http://en.wikipedia.org/w/index.php?oldid=465494912 Contributors: Axiosaurus, Borgx, Brichcja, Calvero JP, Centrx, Dirac66, Jaganath, Kelix, Michael Hardy, Mladjowie, OMCV, Omegakent, Ophiucusthesorceror, Out of Phase User, Passw0rd, Petergans, Physchim62, Puppy8800, Sadads, Smokefoot, T.vanschaik, Themusicking, Tijmz, V8rik, YanA, 33 anonymous edits

Image Sources, Licenses and Contributors

41

Image Sources, Licenses and Contributors


File:HCo(CO)4-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:HCo(CO)4-3D-balls.png License: Public Domain Contributors: Benjah-bmm27 File:Metal-EDTA.svg Source: http://en.wikipedia.org/w/index.php?title=File:Metal-EDTA.svg License: Public Domain Contributors: Smokefoot derivative work: Chamberlain2007 (talk) File:CoA6Cl3.png Source: http://en.wikipedia.org/w/index.php?title=File:CoA6Cl3.png License: Public Domain Contributors: Cwbm (commons), Smokefoot File:CFT - Low Spin Splitting Diagram 2.png Source: http://en.wikipedia.org/w/index.php?title=File:CFT_-_Low_Spin_Splitting_Diagram_2.png License: GNU Free Documentation License Contributors: YanA Image:CFT - High Spin Splitting Diagram 2.png Source: http://en.wikipedia.org/w/index.php?title=File:CFT_-_High_Spin_Splitting_Diagram_2.png License: GNU Free Documentation License Contributors: YanA File:Crystal Field Splitting 4.png Source: http://en.wikipedia.org/w/index.php?title=File:Crystal_Field_Splitting_4.png License: GNU Free Documentation License Contributors: YanA File:Colorwheel.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Colorwheel.jpg License: Public Domain Contributors: Original uploader was Tabatharose at en.wikipedia Image:Octahedral crystal-field splitting.png Source: http://en.wikipedia.org/w/index.php?title=File:Octahedral_crystal-field_splitting.png License: GNU Free Documentation License Contributors: en:YanA File:Pentagonal bipyramidal.png Source: http://en.wikipedia.org/w/index.php?title=File:Pentagonal_bipyramidal.png License: GNU Free Documentation License Contributors: Original uploader was YanA at en.wikipedia (Original text : YanA) Image:Square antiprismatic.png Source: http://en.wikipedia.org/w/index.php?title=File:Square_antiprismatic.png License: GNU Free Documentation License Contributors: Original uploader was YanA at en.wikipedia (Original text : YanA) Image:Square planar.png Source: http://en.wikipedia.org/w/index.php?title=File:Square_planar.png License: GNU Free Documentation License Contributors: Original uploader was YanA at en.wikipedia (Original text : YanA) Image:Square pyramidal.png Source: http://en.wikipedia.org/w/index.php?title=File:Square_pyramidal.png License: GNU Free Documentation License Contributors: Original uploader was YanA at en.wikipedia (Original text : YanA) Image:Tetrahedral.png Source: http://en.wikipedia.org/w/index.php?title=File:Tetrahedral.png License: GNU Free Documentation License Contributors: Original uploader was YanA at en.wikipedia (Original text : YanA) Image:Trigonal bipyramidal.png Source: http://en.wikipedia.org/w/index.php?title=File:Trigonal_bipyramidal.png License: GNU Free Documentation License Contributors: Original uploader was YanA at en.wikipedia (Original text : YanA) Image:Octahedral-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Octahedral-3D-balls.png License: Public Domain Contributors: Benjah-bmm27 Image:AX7E0-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:AX7E0-3D-balls.png License: Public Domain Contributors: Benjah-bmm27 Image:Square-planar-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Square-planar-3D-balls.png License: Public Domain Contributors: Benjah-bmm27, Zzyzx11 Image:Square-pyramidal-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Square-pyramidal-3D-balls.png License: Public Domain Contributors: Benjah-bmm27 Image:Tetrahedral-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Tetrahedral-3D-balls.png License: Public Domain Contributors: Benjah-bmm27 Image:Trigonal-bipyramidal-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Trigonal-bipyramidal-3D-balls.png License: Public Domain Contributors: Benjah-bmm27 Image:Pentagonal-pyramidal-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Pentagonal-pyramidal-3D-balls.png License: Public Domain Contributors: Benjah-bmm27 File:Hexaaquasodium-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Hexaaquasodium-3D-balls.png License: Public Domain Contributors: Ben Mills Image:Oxaliplatin.svg Source: http://en.wikipedia.org/w/index.php?title=File:Oxaliplatin.svg License: Public Domain Contributors: Calvero. Image:Linear2-4Chelate.png Source: http://en.wikipedia.org/w/index.php?title=File:Linear2-4Chelate.png License: Public Domain Contributors: Smokefoot Image:Metal-EDTA.png Source: http://en.wikipedia.org/w/index.php?title=File:Metal-EDTA.png License: Public Domain Contributors: User:Yikrazuul File:M-en1.png Source: http://en.wikipedia.org/w/index.php?title=File:M-en1.png License: Public Domain Contributors: Amada44, Ba10r, DMacks, Pieter Kuiper, Wickey-nl Image:Cu chelate.png Source: http://en.wikipedia.org/w/index.php?title=File:Cu_chelate.png License: Public Domain Contributors: . Original uploader was Petergans at en.wikipedia Image:Equilibrium.svg Source: http://en.wikipedia.org/w/index.php?title=File:Equilibrium.svg License: Public Domain Contributors: L'Aquatique Image:StrikeO.png Source: http://en.wikipedia.org/w/index.php?title=File:StrikeO.png License: Public Domain Contributors: . Original uploader was Petergans at en.wikipedia File:Ferrocene-2D.png Source: http://en.wikipedia.org/w/index.php?title=File:Ferrocene-2D.png License: GNU Free Documentation License Contributors: Original uploader was Benjah-bmm27 at en.wikipedia File:EofRu(bz)2.png Source: http://en.wikipedia.org/w/index.php?title=File:EofRu(bz)2.png License: Public Domain Contributors: Smokefoot File:Eta5-eta3-eta5 Reaction(Colors).png Source: http://en.wikipedia.org/w/index.php?title=File:Eta5-eta3-eta5_Reaction(Colors).png License: Creative Commons Attribution-Sharealike 3.0 Contributors: PedroDaGr8 Image:Trans vs cis ligand.png Source: http://en.wikipedia.org/w/index.php?title=File:Trans_vs_cis_ligand.png License: GNU Free Documentation License Contributors: Zoom321 (talk) Image:LinkageIsomers.png Source: http://en.wikipedia.org/w/index.php?title=File:LinkageIsomers.png License: Public Domain Contributors: Smokefoot Image:Bridgingligand.png Source: http://en.wikipedia.org/w/index.php?title=File:Bridgingligand.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: PedroDaGr8 Image:Di--hydroxo-bis(tetraaquairon(III).png Source: http://en.wikipedia.org/w/index.php?title=File:Di--hydroxo-bis(tetraaquairon(III).png License: Creative Commons Attribution-Sharealike 3.0 Contributors: PedroDaGr8 Image:Mu-Cl.png Source: http://en.wikipedia.org/w/index.php?title=File:Mu-Cl.png License: Public Domain Contributors: Smokefoot Image:DiboraneSchema.png Source: http://en.wikipedia.org/w/index.php?title=File:DiboraneSchema.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: PedroDaGr8 Image:Mu3 compound.png Source: http://en.wikipedia.org/w/index.php?title=File:Mu3_compound.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: PedroDaGr8 Image:MLMBond.png Source: http://en.wikipedia.org/w/index.php?title=File:MLMBond.png License: Public Domain Contributors: OMCV (talk) File:non-innocent1.png Source: http://en.wikipedia.org/w/index.php?title=File:Non-innocent1.png License: Public Domain Contributors: bruintje71 File:Non-innocent2.png Source: http://en.wikipedia.org/w/index.php?title=File:Non-innocent2.png License: Public Domain Contributors: Bruintje71 File:non-innocent3.png Source: http://en.wikipedia.org/w/index.php?title=File:Non-innocent3.png License: Public Domain Contributors: Bruintje71 File:non-innocent4.png Source: http://en.wikipedia.org/w/index.php?title=File:Non-innocent4.png License: Public Domain Contributors: Bruintje71 Image:L-DOPA synthesis2.png Source: http://en.wikipedia.org/w/index.php?title=File:L-DOPA_synthesis2.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: Original uploader was AlChimini at en.wikipedia. Later version(s) were uploaded by Rifleman 82 at en.wikipedia. Image:ChiralLigands.png Source: http://en.wikipedia.org/w/index.php?title=File:ChiralLigands.png License: GNU Free Documentation License Contributors: Original uploader was V8rik at en.wikipedia Image:ChiralLigandsInnerWorkings.png Source: http://en.wikipedia.org/w/index.php?title=File:ChiralLigandsInnerWorkings.png License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Original uploader was V8rik at en.wikipedia Image:Chiralanions.png Source: http://en.wikipedia.org/w/index.php?title=File:Chiralanions.png License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Original uploader was V8rik at en.wikipedia Image:LFTi(III).png Source: http://en.wikipedia.org/w/index.php?title=File:LFTi(III).png License: Public Domain Contributors: Original uploader was Smokefoot at en.wikipedia

License

42

License
Creative Commons Attribution-Share Alike 3.0 Unported //creativecommons.org/licenses/by-sa/3.0/

Vous aimerez peut-être aussi