Jump to content

Palladium dicyanide

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by 134.225.50.57 (talk) at 00:26, 13 January 2011 (Updated to reflect recent publication in Inorg. Chem.). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Palladium(II) cyanides are chemical species with the empirical formula Pd(CN)n(2-n)-. The dicyanide (n = 2, CAS: [2035-66-7]) is a coordination polymer which was the first pure palladium compound isolated. In his attempts to produce pure platinum metal in 1804, W.H. Wollaston added mercuric cyanide to a solution of impure platinum metal in aqua regia to precipitate palladium cyanide which was then ignited to recover palladium metal - a new element. It had long been suspected that the structure of palladium cyanide consists of square planar Pd(II) centers[1] which are linked by bridging cyanide ligands, which are bonded through both the carbon and nitrogen atoms. The CN stretch is seen at 2222 cm-1, typical of bridging cyanide ions. It is now known that the compound commonly known as "Palladium(II) cyanide" is in fact a nanocrystaline material better described using the formula Pd(CN)2.0.29H2O. The interior of the sheets consist of square-planar palladium ions linked by head-to-tail disordered bridging cyanide groups to form nets. These sheets are approximately 3 nm x 3 nm in size and are terminated by an equal number of water and cyanide groups, so that there is no overall charge. The Pd-C and Pd-N bond lengths determined using total neutron diffraction are both 1.98 Å. [2].

The compound is rather insoluble in water with a solubility product of log Ksp = -42.[3]

The slightly more topical palladium(II) cyanide is the dianion [Pd(CN)4]2-. The equilibrium constant for the competition reaction

PdL2+ + 4CN- ⇌ [Pd(CN)4]2- + L, L = 1,4,8,11-tetraazaundecane (2,3,2-tet)[4]

was found to have a value of log K = 14.5.[5] Combination with the formation of the palladium complex with the tetradentate ligand

[Pd(H2O)4]2+ + L ⇌ PdL2+ + 4 H2O, log K = 47.9

gives

[Pd(H2O)4]2+ + 4CN- ⇌ [Pd(CN)4]2- + 4H2O, log β4 = 62.3.

This appears to be the highest formation constant known for any metal ion.[5]

The affinity of Pd for cyanide is so great that palladium metal is attacked by cyanide solutions:

Pd(s) + 2 H+ + 4 CN- ⇌ [Pd(CN)4]2- + H2

This reaction is reminiscent of the “cyanide process” for the extraction of gold, although in the latter reaction O2 is proposed to be involved, to give H2O.[3]

Exchange of between free cyanide ion and [Pd(CN)4]2- has been evaluated by 13C NMR spectroscopy. That exchange occurs at all illustrates the ability of some compounds to be labile (fast reactions) but also stable (high formation constants). The reaction rate is described as follows:

rate = k2[M(CN)42-][CN-], where k2 120 M-1-s-1

The bimolecular kinetics implicate a so-called associative pathway. The associative mechanism of exchange entails rate-limiting attack of cyanide on [Pd(CN)4]2-, possibly with the intermediacy of a highly reactive pentacoordinate species [Pd(CN)5]3-. By comparison, the rate constant for [Ni(CN)4]2- is > 500,000 M-1-s-1, whereas [Pt(CN)4]2-exchanges more slowly at 26 M-1s-1. Such associative reactions are characterized by large negative entropies of activation, in this case: -178 and -143 kJ/(mol·K) for Pd and Pt, respectively.[6]

In organic synthesis palladium cyanide is used in the synthesis of olefinic cyanides from olefins.[7] and as a catalyst in the regioselective reaction between cyanotrimethylsilane and oxiranes .[8]

References

  1. ^ R. B. Janes (1935). "The Diamagnetic Susceptibilities of Palladium Salts". J. Am. Chem. Soc. 57: 471–473. doi:10.1021/ja01306a025.
  2. ^ S. J. Hibble, A. M. Chippindale, E. J. Bilbe, E. Marelli, P. J. F. Harris and A. C. Hannon (2011). "Structures of Pd(CN)2 and Pt(CN)2: Intrinsically Nanocrystaline Materials". Inorg. Chem. 50: 104–113. doi:10.1021/ic101358q.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  3. ^ a b R. D. Hancock, A. Evers (1976). "Formation Constant of Pd(CN)42−". Inorg. Chem. 15: 995–6. doi:10.1021/ic50158a063.
  4. ^ The tetramine 2,3,2-tet, H2N(CH2)2NH(CH2)3NH(CH2)2NH2, is similar to triethylenetetramine (2,2,2-tet) but has an additional methylene group between the two central nitrogen atoms
  5. ^ a b Harrington, James M. (2005). "Determination of formation constants for complexes of very high stability: log β4 for the [Pd(CN)4]2− ion". Inorganica Chimica Acta. 358 (15): 4473–4480. doi:10.1016/j.ica.2005.06.081. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  6. ^ J. J. Pesek, W. R. Mason (1983). "Cyanide Exchange Kinetics for Planar Tetracyanometalate Complexes by Carbon-13 NMR". Inorg. Chem. 22: 2958–2959. doi:10.1021/ic00162a039.
  7. ^ Y. Odaira, T. Oishi, T. Yukawa, S. Tsutsumi (1966). "The Synthesis of Olefinic Cyanides from Olefins by Means of Palladium(II) Cyanide". J. Am. Chem. Soc. 88: 4105–4106. doi:10.1021/ja00969a047.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  8. ^ K. Imi, N. Yanagihara, K. Utimoto (1987). "Reactions of Cyanotrimethylsilane with Oxiranes. Effects of Catalysts or Mediators on Regioselectivity and Ambident Character". J. Org. Chem. 52: 1013–1016. doi:10.1021/jo00382a008.{{cite journal}}: CS1 maint: multiple names: authors list (link)