18-Electron rule

The 18-electron rule is a rule used primarily for predicting formulas for stable metal complexes. [1] The rule rests on the fact that valence shells of a transition metal consists of nine valence orbitals, which collectively can accommodate 18 electrons either as nonbinding electron pairs or as bonding electron pairs. Stated differently, combination of these nine atomic orbitals with ligand orbitals gives rise to nine molecular orbitals that are either metal-ligand bonding or non-bonding. When a metal complex has 18 valence electrons, it is said to have achieved the same electron configuration as the noble gas at the end of the period. The rule and its exceptions are similar to the application of the octet rule to main group elements. The rule is not helpful for complexes of the s-block metals, the lanthanides, and the actinides. The rule was first proposed by American chemist Irving Langmuir in 1921. [2] [3]

Contents

Applicability of the 18-electron rule

Although the majority of metal complexes do not satisfy the 18-electron rule, the rule usefully predicts the formulas for low-spin complexes of the Cr, Mn, Fe, and Co triads. Well-known examples include ferrocene, iron pentacarbonyl, chromium carbonyl, and nickel carbonyl.

Ligands in a complex determine the applicability of the 18-electron rule. In general, complexes that obey the rule are composed at least partly of π-acid ligands. This kind of ligand exerts a very strong ligand field, which lowers the energies of the resultant molecular orbitals and thus favorably occupied. Typical ligands include olefins, phosphines, and CO. Complexes of π-acids typically feature metal in a low-oxidation state. The relationship between oxidation state and the nature of the ligands is rationalized within the framework of π backbonding.

Consequences for reactivity

Compounds that obey the 18 VE rule are typically "exchange inert." Examples include [Co(NH3)5Cl]2+, Mo(CO)6, and [Fe(CN)6]4-. In such cases, in general ligand exchange occurs via dissociative substitution mechanisms, wherein the rate of reaction is determined by the rate of dissociation of a ligand. On the other hand, 18-electron compounds can be highly reactive toward electrophiles such as protons, and such reactions are associative in mechanism, being acid-base reactions.

Complexes with fewer than 18 valence electrons tend to show enhanced reactivity. Thus, the 18-electron rule is often a recipe for non-reactivity in either a stoichiometric or a catalytic sense.

Violations to the 18-electron rule

16e complexes

A popular class of complexes that violate the 18e rule are the 16e complexes with d8 configurations. Examples are especially prevalent for derivatives of the cobalt and nickel triads. Such compounds are typically square-planar. The most famous example is Vaska's complex (IrCl(CO)(PPh3)2), [PtCl4]2−, and Zeise's salt [PtCl32-C2H4)]. In such complexes, the dz2 orbital is doubly occupied and nonbonding.

Many catalytic cycles operate via complexes that alternate between 18e and square-planar 16 configurations. Examples include Monsanto acetic acid synthesis, hydrogenations, hydroformylations, olefin isomerizations, and some alkene polymerizations.

Other violations can be classified according to the kinds of ligands on the metal center.

Bulky ligands

Bulky ligands can preclude the approach of the full complement of ligands that would allow the metal to achieve the 18 electron configuration. Examples:

Sometimes such complexes engage in agostic interactions with the hydrocarbon framework of the bulky ligand. For example:

High-spin complexes

High-spin metal complexes have singly occupied orbitals and may not have any empty orbitals into which ligands could donate electron density. In general, there are few or no π-acidic ligands in the complex. These singly occupied orbitals can combine with the singly occupied orbitals of radical ligands (e.g., oxygen), or addition of a strong field ligand can cause electron-pairing, thus creating a vacant orbital that it can donate into. Examples:

Complexes containing strongly pi-donating ligands often violate the 18-electron rule. These ligands include fluoride (F), oxide (O2−), nitride (N3−), alkoxide (RO), and imide (oxide (RN2−). Examples:

In the latter case, there is substantial donation of the nitrogen lone pairs to the Mo (so the compound could also be described as a 16 VE compound). This can be seen from the short Mo-N bond length, and from the angle Mo - N - C(R), which is nearly 180°. Counter-examples:

In these cases, the M=O bonds are "pure" double bonds (i.e., no donation of the lone pairs of the oxygen to the metal), as reflected in the relatively long bond distances.

Pi-donating ligands

Ligands where the coordinating atom bear nonbonding lone pairs often stabilize unsaturated complexes. Metal amides and alkoxides often violate the 18e rule.

Combinations of effects

The above factors can sometimes combine. Examples include

Higher electron counts

Some complexes have more than 18 electrons. Examples:

Often, cases where complexes have more than 18 valence electrons are attributed to electrostatic forces - the metal attracts ligands to itself to try to counterbalance its positive charge, and the number of electrons it ends up with is unimportant. In the case of the metallocenes, the chelating nature of the cyclopentadienyl ligand stabilizes its bonding to the metal. Somewhat satisfying are the two following observations: (i) cobaltocene is a strong electron donor, readily forming the 18-electron cobaltocenium cation and (ii) nickelocene tends to react with substrates to give 18-electron complexes, e.g. CpNiCl(PR3) and free CpH.

See also

References

  1. ^ Langmuir, I. (1921). "Types of Valence". Science 54 (1386): 59–67. doi:10.1126/science.54.1386.59. 
  2. ^ The Origin of the 18-Electron Rule William B. Jensen Journal of Chemical Education 2005 82 (1), 28 doi:10.1021/ed082p28
  3. ^ Langmuir, I. (1921). "Types of Valence". Science 54 (1386): 59–67. doi:10.1126/science.54.1386.59. 

Further reading