Associative substitution

Source: Wikipedia, the free encyclopedia.

Associative substitution describes a pathway by which compounds interchange ligands. The terminology is typically applied to organometallic and coordination complexes, but resembles the Sn2 mechanism in organic chemistry. The opposite pathway is dissociative substitution, being analogous to the Sn1 pathway. Intermediate pathways exist between the pure associative and pure dissociative pathways, these are called interchange mechanisms.[1][2]

Associative pathways are characterized by

coordinatively unsaturated or contain a ligand that can change its bonding to the metal, e.g. change in hapticity or bending of a nitrogen oxide ligand (NO). In homogeneous catalysis, the associative pathway is desirable because the binding event, and hence the selectivity of the reaction, depends not only on the nature of the metal catalyst but also on the substrate
.

Examples of associative mechanisms are commonly found in the chemistry of 16e

Associative interchange pathway

In many substitution reactions, well-defined intermediates are not observed, when the rate of such processes are influenced by the nature of the entering ligand, the pathway is called associative interchange, abbreviated Ia.[3] Representative is the interchange of bulk and coordinated water in [V(H2O)6]2+. In contrast, the slightly more compact ion [Ni(H2O)6]2+ exchanges water via the Id.[4]

Effects of ion pairing

Polycationic complexes tend to form ion pairs with anions and these ion pairs often undergo reactions via the Ia pathway. The electrostatically held nucleophile can exchange positions with a ligand in the first coordination sphere, resulting in net substitution. An illustrative process comes from the "anation" (reaction with an anion) of chromium(III) hexaaquo complex:

[Cr(H2O)6]3+ + SCN ⇌ {[Cr(H2O)6], NCS}2+
{[Cr(H2O)6], NCS}2+ ⇌ [Cr(H2O)5NCS]2+ + H2O

Special ligand effects

In special situations, some ligands participate in substitution reactions leading to associative pathways. These ligands can adopt multiple motifs for binding to the metal, each of which involves a different number of electrons "donated." A classic case is the

indenyl effect in which an indenyl ligand reversibly "slips' from pentahapto (η5) coordination to trihapto (η3). Other pi-ligands behave in this way, e.g. allyl3 to η1) and naphthalene
6 to η4). Nitric oxide typically binds to metals to make a linear MNO arrangement, wherein the nitrogen oxide is said to donate 3e to the metal. In the course of substitution reactions, the MNO unit can bend, converting the 3e linear NO ligand to a 1e bent NO ligand.

SN1cB mechanism

The rate for the

SN1cB mechanism
.

Eigen-Wilkins mechanism

The Eigen-Wilkins mechanism, named after chemists Manfred Eigen and R. G. Wilkins,[5] is a mechanism and rate law in coordination chemistry governing associative substitution reactions of octahedral complexes. It was discovered for substitution by ammonia of a chromium-(III) hexaaqua complex.[6][7] The key feature of the mechanism is an initial rate-determining pre-equilibrium to form an encounter complex ML6-Y from reactant ML6 and incoming ligand Y. This equilibrium is represented by the constant KE:

ML6 + Y ⇌ ML6-Y

The subsequent dissociation to form product is governed by a rate constant k:

ML6-Y → ML5Y + L

A simple derivation of the Eigen-Wilkins rate law follows:[8]

[ML6-Y] = KE[ML6][Y]
[ML6-Y] = [M]tot - [ML6]
rate = k[ML6-Y]
rate = kKE[Y][ML6]

Leading to the final form of the rate law, using the steady-state approximation (d[ML6-Y] / dt = 0),

rate = kKE[Y][M]tot / (1 + KE[Y])

Eigen-Fuoss equation

A further insight into the pre-equilibrium step and its equilibrium constant KE comes from the Fuoss-Eigen equation proposed independently by Eigen and R. M. Fuoss:

KE = (4πa3/3000) x NAexp(-V/RT)

Where a represents the minimum distance of approach between complex and ligand in solution (in cm), NA is the

electrostatic potential energy
of the ions at that distance:

V = z1z2e2/4πaε

Where z is the charge number of each species and ε is the vacuum permittivity.

A typical value for KE is 0.0202 dm3mol−1 for neutral particles at a distance of 200 pm.[9] The result of the rate law is that at high concentrations of Y, the rate approximates k[M]tot while at low concentrations the result is kKE[M]tot[Y]. The Eigen-Fuoss equation shows that higher values of KE (and thus a faster pre-equilibrium) are obtained for large, oppositely-charged ions in solution.

References

  1. .
  2. .
  3. .
  4. .
  5. ^ M. Eigen, R. G. Wilkins: Mechanisms of Inorganic Reactions. In: Advances in Chemistry Series. Nr. 49, 1965, S. 55. American Chemical Society, Washington, D. C.
  6. .
  7. ^ Atkins, P. W. (2006). Shriver & Atkins inorganic chemistry. 4th ed. Oxford: Oxford University Press