Metal aquo complex

Metal aquo complexes are coordination compounds containing metal ions with only water as a ligand. These complexes are the predominant species in aqueous solutions of many metal salts, such as metal nitrates, sulfates, and perchlorates. They have the general stoichiometry [M(H2O)n]z+. Their behavior underpins many aspects of environmental, biological, and industrial chemistry. This article focuses on complexes where water is the only ligand ("homoleptic aquo complexes"), but of course many complexes are known to consist of a mix of aquo and other ligands.[1]


Stoichiometry and structure

Structure of an octahedral metal aquo complex.
Chromium(II) ion in aqueous solution.

The majority of aquo-complexes have the general formula [M(H2O)6]n+, with n= 2 or 3; they have an octahedral structure. Some examples are listed in the following table.

Complexcolourelectron config.M–O distance (Å)[2]water exchange
rate (s−1, 25 °C)[3]
M2+/3+ self-exchange
rate (M−1s−1, 25 °C)
[Ti(H2O)6]3+ violet (t2g)1 2.025 1.8 x 105 n.a.
[V(H2O)6]2+ violet (t2g)3 2.12 fast
[V(H2O)6]3+ yellow (t2g)2 1.991[4] 5.0 x 102 fast
[Cr(H2O)6]2+ blue (t2g)3(eg)1 2.06, 2.33 1.2 x 108 slow
[Cr(H2O)6]3+ violet (t2g)3 1.961 2.4 x 10−6 slow
[Mn(H2O)6]2+ pale pink (t2g)3(eg)2 2.177 2.1 x 107 n.a.
[Fe(H2O)6]2+ blue-green (t2g)4(eg)2 2.095 4.4 x 106 fast
[Fe(H2O)6]3+ pale violet (t2g)3(eg)2 1.990 1.6 x 102 fast
[Co(H2O)6]2+ pink (t2g)5(eg)2 2.08 3.2 x 106 n.a.
[Ni(H2O)6]2+ green (t2g)6(eg)2 2.05 3.2 x 104 n.a.
[Cu(H2O)6]2+ blue (t2g)6(eg)3 1.97, 2.30 5.7 x 109 n.a.

Tutton's salts, are crystalline compounds with the generic formula (NH4)2M(SO4)2.(H2O)6 (M=V2+, Cr2+, Mn2+, Co2+, Ni2+ and Cu2+). Alums, MM'(SO4)2(H2O)12, are also double salts. Both sets of salts contain hexa-aquo metal cations.

Silver(I) forms [Ag(H2O)4]+, a rare example of a tetrahedral aquo complex.[5] Palladium(II) and platinum(II) form square planar aquo complexes with the stoichiometry [M(H2O)4]2+. Aquo complexes of lanthanide(III) ions are eight- and nine-coordinate, reflecting the large size of the metal centres. Complexes such as [Mo2(H2O)8]4+ and [Rh2(H2O)10]4+ contain metal-metal bonds.[5]

Monomeric aquo complexes of Zr, Hf, Nb, Ta, Mo, W, Mn, Tc, Re and Os in oxidation states 4-7 cannot be detected. [6] The ion [M(H2O)n]m+ is subject to extensive hydrolysis. For example, [Ti(H2O)6]4+ is unknown: the hydrolyzed species [Ti(OH)2(H2O)n]2+is the principal species in dilute solutions. [7] With the higher oxidation states the effective electrical charge on the cation is further reduced by the formation of oxo-complexes. For example vanadium(V) forms vanadyl complexes. The hypothetical reaction

[V(H2O)6]5+ → [VO(H2O)5]3+ + 2H+

goes to completion and the hexa-aquo ion cannot be detected in solutions of vanadium(V) compounds. With chromium(VI) and manganese (VII) only oxyanions are known.

Reactions

Several reactions are fundamental to the behavior of metal aquo ions: ligand exchange, electron-transfer, and acid-base reactions of the O-H bonds.

Water exchange

Ligand exchange involve replacement of a water ligand ("coordinated water") with water in solution ("bulk water"). Often the process is represented using labeled water H2O*:

[M(H2O)n]z+ + H2O* → [M(H2O)n-1(H2O*)]z+ + H2O

In the absence of isotopic labeling, the reaction is degenerate, meaning that the free energy change is zero. Rates vary over many orders of magnitude. The main factor affecting rates is charge: highly charged metal aquo cations exchange their water more slowly than singly charged species. Thus, the exchange rates for [Na(H2O)6]+ and [Al(H2O)6]3+ differ by a factor of 109. Electron configuration is also a major factor, illustrated by the fact that the rates of water exchange for [Al(H2O)6]3+ and [Ir(H2O)6]3+ differ by a factor of 109 also.[3] Water exchange usually follows a dissociative substitution pathway, so the rate constants indicate first order reactions.

Electron exchange

This reaction usually applies to the interconversion of di- and trivalent metal ions, which involves the exchange of only one electron. The process is called self-exchange, meaning that the ion appears to exchange electrons with itself. The standard electrode potential for the equilibrium

[M(H2O)6]2+ + [M(H2O)6]3+ [M(H2O)6]3+ + [M(H2O)6]2+
Standard redox potential for the couple M2+, M3+ / V
VCrMnFeCo
-0.26-0.41+1.51+0.77+1.82

shows the increasing stability of the lower oxidation state as atomic number increases. The very large value for the manganese couple is a consequence of the fact that octahedral manganese(II) has zero crystal field stabilization energy (CFSE) but manganese(III) has 3 units of CFSE.[8]

Using labels to keep track of the metals, the self-exchange process is written as:

[M(H2O)6]2+ + [M*(H2O)6]3+ → [M*(H2O)6]2+ + [M(H2O)6]3+

The rates of electron exchange vary widely, the variations being attributable to differing reorganization energies: when the 2+ and 3+ ions differ widely in structure, the rates tend to be slow.[9] The electron transfer reaction proceeds via an outer sphere electron transfer. Most often large reorganizational energies are associated with changes in the population of the eg level, at least for octahedral complexes.

Acid–base reactions

Solutions of metal aquo complexes are acidic owing to the ionization of protons from the water ligands. In dilute solution chromium(III) aquo complex has a pKa of about 4.3:

[Cr(H2O)6]3+ [Cr(H2O)5(OH)]2+ + H+

Thus, the aquo ion is a weak acid, of comparable strength to acetic acid (pKa of about 4.8). This is typical of the trivalent ions. The influence of the electronic configuration on acidity is shown by the fact that [Ru(H2O)6]3+ (pKa = 2.7) is more acidic than [Rh(H2O)6]3+ (pKa =4), despite the fact that Rh(III) is expected to be more electronegative. This effect is related to the stabilization of the pi-donor hydroxide ligand by the (t2g)5 Ru(III) centre.[5]

In more concentrated solutions, some metal hydroxo complexes undergo condensation reactions, known as olation, to form polymeric species. Many minerals, form via olation. Aquo ions of divalent metal ions are less acidic than those of trivalent cations.

The hydrolyzed species often exhibit very different properties from the precursor hexa-aquo complex. For example, water exchange in [Al(H2O)5OH]2+ is some 20000 times faster than in [Al(H2O)6]3+.

References

  1. Mark I. Ogden and Paul D. Beer "Water & O-Donor Ligands" in Encyclopedia of Inorganic Chemistry, Wiley-VCH, 2006, Weinheim. doi:10.1002/0470862106.ia255
  2. For Mn(II), Fe(II), Fe(III): T. K. Sham, J. B. Hastings, M. L. Perlman "Structure and dynamic behavior of transition-metal ions in aqueous solution: an EXAFS study of electron-exchange reactions" J. Am. Chem. Soc., 1980, 102 (18), pp 5904–5906. doi:10.1021/ja00538a033. For Ti(III), V(III), Cr(III): B. Kallies, R. Meier "Electronic Structure of 3d [M(H2O)6]3+ Ions from ScIII to FeIII: A Quantum Mechanical Study Based on DFT Computations and Natural Bond Orbital Analyses" Inorg. Chem., 2001, 40 (13), pp 3101–3112. doi:10.1021/ic001258t
  3. Lothar Helm, André E. Merbach "Inorganic and Bioinorganic Solvent Exchange Mechanisms" Chemical Reviews 2005, volume 105, 1923-1959. doi:10.1021/cr030726o
  4. Cotton, F. A.; Fair, C. K.; Lewis, G. E.; Mott, G. N.; Ross, F. K.; Schultz, A. J.; Williams, J. M. (1984). "Precise Structural Characterizations of the Hexaaquovanadium(III) and Diaquohydrogen Ions. X-ray and Neutron Diffraction Studies of [V(H2O)6][H5O2](CF3SO3)4". Journal of the American Chemical Society. 106: 5319–5323. doi:10.1021/ja00330a047.CS1 maint: uses authors parameter (link)
  5. S. F. Lincoln, D. T. Richens, A. G. Sykes "Metal Aqua Ions" Comprehensive Coordination Chemistry II Volume 1, Pages 515-555. doi:10.1016/B0-08-043748-6/01055-0
  6. An ion such as [Zr(H2O)6]4+ must exist in aqueous solution, but its concentration will be well below the detection limit
  7. Baes, C.F.; Mesmer, R.E. The Hydrolysis of Cations, (1976), Wiley, New York
  8. Burgess, John (1978). Metal Ions in Solution. Chichester: Ellis Horwood. ISBN 0-85312-027-7. p. 236.
  9. R. G. Wilkins Kinetics and Mechanism of Reactions of Transition Metal Complexes, 2nd Edition, VCH, Weinheim, 1991. ISBN 1-56081-125-0

See also

This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.