Cis effect

Last updated

In inorganic chemistry, the cis effect is defined as the labilization (or destabilization) of CO ligands that are cis to other ligands. CO is a well-known strong pi-accepting ligand in organometallic chemistry that will labilize in the cis position when adjacent to ligands due to steric and electronic effects. The system most often studied for the cis effect is an octahedral complex M(CO)
5
X
where X is the ligand that will labilize a CO ligand cis to it. Unlike the trans effect, which is most often observed in 4-coordinate square planar complexes, the cis effect is observed in 6-coordinate octahedral transition metal complexes. It has been determined that ligands that are weak sigma donors and non-pi acceptors seem to have the strongest cis-labilizing effects. Therefore, the cis effect has the opposite trend of the trans-effect, which effectively labilizes ligands that are trans to strong pi accepting and sigma donating ligands. [1] [2] [3]

Contents

Electron counting in metal carbonyl complexes

Group 6 and group 7 transition metal complexes M(CO)
5
X
have been found to be the most prominent in regards to dissociation of the CO cis to ligand X. [4] CO is a neutral ligand that donates 2 electrons to the complex, and therefore lacks anionic or cationic properties that would affect the electron count of the complex. For transition metal complexes that have the formula M(CO)
5
X
, group 6 metals (M0, where the oxidation state of the metal is zero) paired with neutral ligand X, and group 7 metals (M+, where the oxidation state of the metal is +1), paired anionic ligands, will create very stable 18 electron complexes. Transition metal complexes have 9 valence orbitals, and 18 electrons will in turn fill these valences shells, creating a very stable complex, which satisfies the 18-electron rule. The cis-labilization of 18 e complexes suggests that dissociation of ligand X in the cis position creates a square pyramidal transition state, which lowers the energy of the M(CO)
4
X
complex, enhancing the rate of reaction. [5] The scheme below shows the dissociation pathway of a CO ligand in the cis and trans position to the X, followed by the association of ligand Y. This is an example of a dissociative mechanism, where an 18 e complex loses a CO ligand, making a 16 e intermediate, and a final complex of 18 e results from an incoming ligand inserting in place of the CO. This mechanism resembles the SN1 mechanism in organic chemistry, and applies to coordination compounds as well. [6]

Wiki figure.png

Figure 1. Intermediates in the substitution of M(CO)
5
X
complexes. If ligands X and Y are neutral donors to the complex:

M = Group 6 metal (m = 0)

M = Group 7 metal (m = +1)

Ligands effects on CO cis-labilization

The order of ligands which possess cis-labilizing effects are as follows: CO, AuPPh3, H, SnPh3, GePh3, M(CO)n < P(O)Ph3 < PPh3 < I < CH3SO2, NC5H5 < CH3CO < Br, NCO < Cl < NO3

Anionic ligands such as F, Cl, OH, and SH have particularly strong CO labilizing effects in [M(CO)
5
L]
complexes. This is because these ligands will stabilize the 16 e intermediate by electron donation from the p-pi lone pair donor orbital. [7] Other sulfur-containing ligands, particularly thiobenzoate, are other examples of particularly useful CO cis-labilizing ligands, which can be explained by stabilization of the intermediate that results upon CO dissociation. This can be attributed to the partial interaction of the oxygen from the thiobenzoate and the metal, which can eliminate solvent effects that can occur during ligand dissociation in transition metal complexes. [4]

Note that the strongest labilizing effects come from ligands that are weak sigma donors with virtually no pi-accepting behavior. The cis effect can be attributed to the role of ligand X in stabilizing the transition state. It has also been determined that labilizing X ligands do in fact strengthen the M-CO bond trans to X, which is hypothesized to be due to the weak pi-accepting and/or sigma donating behavior of ligand X. This lack of strong sigma donation/pi-accepting will allow the CO (a strong pi-acceptor) trans to ligand X to pull electron density toward it, strengthening the M-CO bond. This phenomenon is further supported by the evidence from extensive studies on the trans effect, which in turn shows how ligands that are actually strong sigma donors and pi-acceptors weaken the M-L bond trans to them. Since the cis and trans effects seem to have generally opposite trends, the electronic argument supports both phenomena. Further evidence for cis labilization of CO can be attributed to the CO ligands being in competition for the dxy, dyz, and dxz orbitals. This argument especially holds true when the X is a halogen. [8]

Related Research Articles

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

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

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

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

<span class="mw-page-title-main">Organometallic chemistry</span> Study of organic compounds containing metal(s)

Organometallic chemistry is the study of organometallic compounds, chemical compounds containing at least one chemical bond between a carbon atom of an organic molecule and a metal, including alkali, alkaline earth, and transition metals, and sometimes broadened to include metalloids like boron, silicon, and selenium, as well. Aside from bonds to organyl fragments or molecules, bonds to 'inorganic' carbon, like carbon monoxide, cyanide, or carbide, are generally considered to be organometallic as well. Some related compounds such as transition metal hydrides and metal phosphine complexes are often included in discussions of organometallic compounds, though strictly speaking, they are not necessarily organometallic. The related but distinct term "metalorganic compound" refers to metal-containing compounds lacking direct metal-carbon bonds but which contain organic ligands. Metal β-diketonates, alkoxides, dialkylamides, and metal phosphine complexes are representative members of this class. The field of organometallic chemistry combines aspects of traditional inorganic and organic chemistry.

A substitution reaction is a chemical reaction during which one functional group in a chemical compound is replaced by another functional group. Substitution reactions are of prime importance in organic chemistry. Substitution reactions in organic chemistry are classified either as electrophilic or nucleophilic depending upon the reagent involved, whether a reactive intermediate involved in the reaction is a carbocation, a carbanion or a free radical, and whether the substrate is aliphatic or aromatic. Detailed understanding of a reaction type helps to predict the product outcome in a reaction. It also is helpful for optimizing a reaction with regard to variables such as temperature and choice of solvent.

The Stille reaction is a chemical reaction widely used in organic synthesis. The reaction involves the coupling of two organic groups, one of which is carried as an organotin compound (also known as organostannanes). A variety of organic electrophiles provide the other coupling partner. The Stille reaction is one of many palladium-catalyzed coupling reactions.

<span class="mw-page-title-main">Pi backbonding</span> Form of interaction between two atoms

In chemistry, π back bonding is a π-bonding interaction between a filled (or half filled) orbital on one atom and a vacant orbital on an adjacent atom. This type of interaction stabilizes electron rich atoms by allowing them to dissipate charge into neighboring atoms.. It is especially common in the organometallic chemistry of low-valent transition metals with ligands such as carbon monoxide, olefins, or phosphines. Electrons from the metal are used to bond to the ligand, dissipating excess negative charge. Compounds where π back bonding is prominent include Ni(CO)4, Zeise's salt, and molybdenym and iron dinitrogen complexes.

Reductive elimination is an elementary step in organometallic chemistry in which the oxidation state of the metal center decreases while forming a new covalent bond between two ligands. It is the microscopic reverse of oxidative addition, and is often the product-forming step in many catalytic processes. Since oxidative addition and reductive elimination are reverse reactions, the same mechanisms apply for both processes, and the product equilibrium depends on the thermodynamics of both directions.

Oxidative addition and reductive elimination are two important and related classes of reactions in organometallic chemistry. Oxidative addition is a process that increases both the oxidation state and coordination number of a metal centre. Oxidative addition is often a step in catalytic cycles, in conjunction with its reverse reaction, reductive elimination.

beta-Hydride elimination

β-Hydride elimination is a reaction in which an alkyl group bonded to a metal centre is converted into the corresponding metal-bonded hydride and an alkene. The alkyl must have hydrogens on the β-carbon. For instance butyl groups can undergo this reaction but methyl groups cannot. The metal complex must have an empty site cis to the alkyl group for this reaction to occur. Moreover, for facile cleavage of the C–H bond, a d electron pair is needed for donation into the σ* orbital of the C–H bond. Thus, d0 metals alkyls are generally more stable to β-hydride elimination than d2 and higher metal alkyls and may form isolable agostic complexes, even if an empty coordination site is available.

In inorganic chemistry, the trans effect is the increased lability of ligands that are trans to certain other ligands, which can thus be regarded as trans-directing ligands. It is attributed to electronic effects and it is most notable in square planar complexes, although it can also be observed for octahedral complexes. The analogous cis effect is most often observed in octahedral transition metal complexes.

<span class="mw-page-title-main">Dimanganese decacarbonyl</span> Chemical compound

Dimanganese decacarbonyl, which has the chemical formula Mn2(CO)10, is a binary bimetallic carbonyl complex centered around the first row transition metal manganese. The first reported synthesis of Mn2(CO)10 was in 1954 at Linde Air Products Company and was performed by Brimm, Lynch, and Sesny. Their hypothesis about, and synthesis of, dimanganese decacarbonyl was fundamentally guided by the previously known dirhenium decacarbonyl (Re2(CO)10), the heavy atom analogue of Mn2(CO)10. Since its first synthesis, Mn2(CO)10 has been use sparingly as a reagent in the synthesis of other chemical species, but has found the most use as a simple system on which to study fundamental chemical and physical phenomena, most notably, the metal-metal bond. Dimanganese decacarbonyl is also used as a classic example to reinforce fundamental topics in organometallic chemistry like d-electron count, the 18-electron rule, oxidation state, valency, and the isolobal analogy.

Transition metal hydrides are chemical compounds containing a transition metal bonded to hydrogen. Most transition metals form hydride complexes and some are significant in various catalytic and synthetic reactions. The term "hydride" is used loosely: some of them are acidic (e.g., H2Fe(CO)4), whereas some others are hydridic, having H-like character (e.g., ZnH2).

In chemistry, dissociative substitution describes a reaction pathway by which compounds interchange ligands. The term is typically applied to coordination and organometallic complexes, but resembles the SN1 mechanism in organic chemistry. This pathway can be well described by the cis effect, or the labilization of CO ligands in the cis position. The opposite pathway is associative substitution, being analogous to SN2 pathway. Pathways that are intermediate between the pure dissociative and pure associative pathways are called interchange mechanisms.

<span class="mw-page-title-main">Organorhodium chemistry</span> Field of study

Organorhodium chemistry is the chemistry of organometallic compounds containing a rhodium-carbon chemical bond, and the study of rhodium and rhodium compounds as catalysts in organic reactions.

<span class="mw-page-title-main">Metal-phosphine complex</span>

A metal-phosphine complex is a coordination complex containing one or more phosphine ligands. Almost always, the phosphine is an organophosphine of the type R3P (R = alkyl, aryl). Metal phosphine complexes are useful in homogeneous catalysis. Prominent examples of metal phosphine complexes include Wilkinson's catalyst (Rh(PPh3)3Cl), Grubbs' catalyst, and tetrakis(triphenylphosphine)palladium(0).

The covalent bond classification (CBC) method, also referred to as LXZ notation, is a way of describing covalent compounds such as organometallic complexes in a way that is not prone to limitations resulting from the definition of oxidation state. Instead of simply assigning a charge to an atom in the molecule, the covalent bond classification method analyzes the nature of the ligands surrounding the atom of interest. According to this method, the interactions that allow for coordination of the ligand can be classified according to whether it donates two, one, or zero electrons. These three classes of ligands are respectively given the symbols L, X, and Z. The method was published by Malcolm L. H. Green in 1995.

In covalent bond classification, a Z-type ligand refers to a ligand that accepts two electrons from the metal center. This is in contrast to X-type ligands, which form a bond with the ligand and metal center each donating one electron, and L-type ligands, which form a bond with the ligand donating two electrons. Typically, these Z-type ligands are Lewis acids, or electron acceptors. They are also known as zero-electron reagents.

In organometallic chemistry, a transition metal alkyne complex is a coordination compound containing one or more alkyne ligands. Such compounds are intermediates in many catalytic reactions that convert alkynes to other organic products, e.g. hydrogenation and trimerization.

<span class="mw-page-title-main">Transition-metal allyl complex</span>

Transition-metal allyl complexes are coordination complexes with allyl and its derivatives as ligands. Allyl is the radical with the connectivity CH2CHCH2, although as a ligand it is usually viewed as an allyl anion CH2=CH−CH2, which is usually described as two equivalent resonance structures.

In chemistry, transition metal silyl complexes describe coordination complexes in which a transition metal is bonded to an anionic silyl ligand, forming a metal-silicon sigma bond. This class of complexes are numerous and some are technologically significant as intermediates in hydrosilylation. These complexes are a subset of organosilicon compounds.

References

  1. Miessler, Gary O. Spessard, Gary L. (2010). Organometallic chemistry (2nd ed.). New York: Oxford University Press. ISBN   978-0195330991.{{cite book}}: CS1 maint: multiple names: authors list (link)
  2. Atwood, Jim D. (1997). Inorganic and organometallic reaction mechanisms (2. ed.). New York [u.a.]: Wiley-VCH. ISBN   978-0471188971.
  3. Atkins, Peter (2010). Shriver & Atkins' inorganic chemistry (5th ed.). New York: W. H. Freeman and Co. ISBN   978-1429218207.
  4. 1 2 Atwood, J.; Brown, Theodore L. (1976). "Cis labilization of ligand dissociation. 3. Survey of group 6 and 7 six-coordinate carbonyl compounds. The site preference model for ligand labilization effects". J. Am. Chem. Soc. 98 (11): 3160–3166. doi:10.1021/ja00427a017.
  5. Jensen, W. (2005). "The Origin of the 18-Electron Rule". J. Chem. Educ. 82 (1): 28. Bibcode:2005JChEd..82...28J. doi:10.1021/ed082p28.
  6. Hill AF, Fink MJ (2010). Advances in Organometallic Chemistry. Oxford: Academic Pr. ISBN   978-0-12-378649-4.
  7. Kovacs, A.; Frenking, Gernot (2001). "Stability and Bonding Situation of Electron-Deficient Transition-Metal Complexes. Theoretical Study of the CO-Labilizing Effect of Ligands L in [W(CO)5L] (L = C2H2, NCH, N2, C2H4, OH2, SH2, NH3, F, Cl, OH, SH) and [W(CO)4L]2− (L2− = O2C2H22−, S2C2H22−) and the Structure of the 16-Valence-Electron Complexes [W(CO)4L] and [W(CO)3L]2−". Organometallics. 20 (12): 2510–2524. doi:10.1021/om0101893.
  8. Asali, K. J.; Janaydeh, Husam Al (2003). "Transition Metal Chemistry". Transit. Met. Chem. 28 (2): 193–198. doi:10.1023/A:1022953903025. S2CID   91996293.