You are on page 1of 5

Wacker process

From Wikipedia, the free encyclopedia

Jump to: navigation, search The Wacker process or the Hoechst-Wacker process (named after the chemical companies of the same name) originally referred to the oxidation of ethylene to acetaldehyde by oxygen in water in the presence of a tetrachloropalladate catalyst.[1] The same basic reaction is currently used to produce aldehydes and ketones from a number of alkenes with the Monsanto process for producing acetic acid. This chemical reaction, a German invention, was the first organometallic and organopalladium reaction applied on an industrial scale. The Wacker process is similar to hydroformylation, which is also an industrial process and also leads to aldehyde compounds. The differences are that hydroformylation promotes chain extension, and uses a rhodium-based catalyst system. The Wacker process is an example of homogeneous catalysis. The palladium complex with ethylene is reminiscent of Zeise's salt, K[PtCl3(C2H4)] which is a heterogeneous catalyst.

Contents
[hide]

1 History 2 Reaction mechanism 3 Mechanism summary 4 Wacker-Tsuji oxidation 5 References

[edit] History
The development of the chemical process now known as the Wacker process began in 1956 at Wacker Chemie.[2] At the time, many simple aliphatic compounds were produced from acetylene (as calcium carbide) but the construction of a new oil refinery in Cologne by Esso close to a Wacker site, combined with the realization that ethylene would be a cheaper raw-material prompted Wacker to investigate its potential uses. As part of the ensuing research effort, a reaction of ethylene and oxygen over palladium on carbon in a quest for ethylene oxide unexpectedly gave evidence for the formation of acetaldehyde (simply based on smell). More research into this ethylene to acetaldehyde conversion resulted in 1957 in a gas-phase reaction patent using a heterogeneous catalyst.[3] In the meanwhile Hoechst AG joined the race and after a patent filing forced Wacker into a partnership called Aldehyd GmbH. The heterogeneous process ultimately failed due to catalyst inactivation and was replaced by the water-based homogeneous system for which a pilot plant was operationally in 1958. Problems with the aggressive catalyst solution were solved by adopting titanium (newly available for industrial use) as construction material for reactors and pumps. Production plants went into operation in 1960.

[edit] Reaction mechanism


The reaction mechanism for the industrial Wacker process (olefin oxidation via palladium(II) chloride) has received significant attention for several decades, and parts of it are still a contentious subject. A modern formulation is described below:

The catalytic cycle can also be described as follows: [PdCl4]2 + C2H4 + H2O CH3CHO + Pd + 2 HCl + 2 Cl Pd + 2 CuCl2 + 2 Cl [PdCl4]2 + 2 CuCl 2 CuCl + O2 + 2 HCl 2 CuCl2 + H2O Note that all catalysts are regenerated and only the alkene and oxygen are consumed. Without copper(II) chloride as an oxidizing agent Pd(0) metal (resulting from reductive elimination of Pd(II) in the final step) would precipitate out and the reaction would come to a halt (the stoichiometric reaction without catalyst regeneration was discovered in 1894). Air, pure oxygen, or a number of other oxidizers can then oxidise the resultant CuCl back to CuCl2, allowing the cycle to repeat. The initial stoichiometric reaction was first reported by Phillips[4][5] and the Wacker reaction was first reported by Smidt et al.[6][7][8]

[edit] Mechanism summary

Early mechanistic studies from the 1960s elucidated much of the oxidation process.[9][10] Several interesting key points were found, and these observations are generally not considered controversial: 1. There is no H/D exchange seen in this reaction. Experiments using C2D4 in water generate CD3CDO, and runs with C2H4 in D2O generate CH3CHO. Thus, keto-enol tautomerization is not a possible mechanistic step. 2. There is a negligible kinetic isotope effect with fully deuterated reactants (k H/k D=1.07). Hence, it is inferred that hydride transfer is not a rate-determining step. 3. A significant competitive isotope effect with C2H2D2, (k H/k D= ~1.9), suggests that the rate determining step should be prior to oxidized product formation. 4. Running the reaction under higher concentrations of chloride and copper(II) chloride result in the formation of a new product, chlorohydrin. Based on these observations, it is generally accepted the rate-determining step occurs before a series of hydride rearrangements. Many experimental and theoretical investigations have sought to identify the rate determining step, but a unifying interpretation of these experiments has been difficult to attain. Most mechanistic studies on the Wacker process have focused on identifying whether nucleophilic attack occurred via an external (anti-addition) pathway or via an internal (synaddition) pathway. Using arguments supported by kinetics experiments, Henry inferred the mechanism for nucleophilic attack would be by an internal (syn-) pathway.[11] Later, stereochemical studies by Stille and coworkers[12][13][14] yielded chemical products that indicated the Wacker process proceeds via anti-addition; however, since these experiments were run under conditions significantly different from industrial Wacker process conditions, the conclusions were disputed. Contemporary stereochemical studies using normal industrial Wacker conditions (except with high chloride and high copper chloride concentrations) also yielded products that inferred nucleophilic attack was an anti-addition reaction.[15] The published results of these two independent stereochemical studies were accepted by the chemistry community as proof that the standard reaction process occurs via an anti-addition step, and many reference texts have not renewed their presentation of the Wacker process mechanism since this point. Later kinetic studies by Henry and coworkers used isotopically substituted allyl alcohols at standard industrial conditions (with low-chloride concentrations) to probe the reaction mechanisms.[16][17] Those results showed nucleophilic attack is a slow process, while the proposed mechanisms explaining the earlier stereochemical studies assumed nucleophilic attack to be a fast process. Subsequent stereochemical studies by Patrick M. Henry and coworkers observed different stereoisomers of products that indicated that both pathways occur and are dependent on chloride concentrations.[18][19] However, these studies too are disputed since allyl-alcohols may be sensitive to isomerization reactions, and different stereoisomers may be formed from those reactions and not from the standard Wacker process.

In summary, experimental evidence seems to support that syn-addition occurs under lowchloride reaction concentrations (< 1 mol/L, industrial process conditions), while anti-addition occurs under high-chloride (> 3mol/L) reaction concentrations, probably due to chloride ions saturating the catalyst and inhibiting the inner-sphere mechanism. However, the exact pathway and the reason for this switching of pathways is still unknown. Further complicating the Wacker process mechanism is questions about the role of copper chloride in the mechanism. Most mechanistic theories so far presented have invoked Occam's razor, and assumed copper does not play a role in the olefin oxidation mechanisms. Yet, experiments by Stangl and Jira[20] found chlorohydrin formation was dependent on copper chloride concentrations. Work by Hosokawa and coworkers[21] yielded a crystallized product containing copper chloride, indicating it may have a non-innocent role in olefin oxidation. Finally, an ab initio studio by Comas-Vives, et. al. [22] concluded that the anti-addition was the preferred one under low chloride and copper co-catalyst concentrations. This was afterwards confirmed by experiments by Anderson and Sigman[23] which yielded a different kinetic rate law for olefin oxidation without copper than reactions using industrial conditions. While these works complicate the picture of the Wacker process mechanism, one should probably infer that this and related chemistry can be sensitive to reaction conditions, and multiple different reaction pathways may be in play. Another key step in the Wacker process is the migration of the hydrogen from oxygen to chlorine and formation of the C-O double bond. This step is often regarded to proceed through a so-called -hydride elimination with a four-membered cyclic transition state:

In silico studies[24][25] argue that the transition state for this reaction step is unfavorable and an alternative reductive elimination reaction mechanism is in play. The proposed reaction steps are likely assisted by water molecule in solution acting as a catalyst.

[edit] Wacker-Tsuji oxidation


The so-called Wacker-Tsuji oxidation is the laboratory scale version of the above reaction, for example the conversion of 1-decene to 2-decanone with palladium(II) chloride and copper(II) chloride in a water / dimethylformamide solvent mixture in the presence of air:[26]

Some of the related chemistry on Wacker-like oxidations and aminations have been reviewed by Stahl and coworkers.[27]

[edit] References
1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. ^ Translated in part from de:Wacker-Verfahren. ^ Acetaldehyde from Ethylene A Retrospective on the Discovery of the Wacker Process Reinhard Jira Angew. Chem. Int. Ed. 2009, 48, 9034-9037 doi:10.1002/anie.200903992 ^ J. Smidt, W. Hafner, J. Sedlmeier, R. Jira, R. Rottinger (Cons. f.elektrochem.Ind.), DE 1 049 845, 1959, Anm. 04.01.1957. ^ F.C. Phillips, Am. Chem. J., 1894, 16, 255-277. ^ F.C. Phillips, Z. Anorg. Chem., 1894, 6, 213-228. ^ J. Smidt, W. Hafner, R. Jira, J. Sedlmeier, R. Sieber, R. Rttinger, and H. Kojer, Angew. Chem., 1959, 71, 176-182. doi:10.1002/ange.19590710503 ^ W. Hafner, R. Jira, J. Sedlmeier, and J. Smidt, Chem. Ber., 1962, 95, 1575-1581. ^ J. Smidt, W. Hafner, R. Jira, R. Sieber, J. Sedlmeier, and A. Sabel, Angew. Chem. Int. Ed. Engl., 1962, 1, 80-88. ^ Henry, Patrick M. In Handbook of Organopalladium Chemistry for Organic Synthesis; Negishi, E., Ed.; Wiley & Sons: New York, 2002; p 2119. ISBN 0471315060 ^ J. A. Keith and P. M. Henry, Angew. Chem. Int. Ed. 2009, 48, 9038-9049doi:10.1002/anie.200902194 ^ P.M. Henry, J. Am. Chem. Soc., 1964, 86, 3246-3250. ^ James, D.E., Stille, J.K. J. Organomet. Chem., 1976, 108, 401. doi:10.1021/ja00423a028 ^ Stille, J.K., Divakarumi, R.J., J. Organomet. Chem., 1979, 169, 239; ^ James, D.E., Hines, L.F., Stille, J.K. J. Am. Chem. Soc., 1976, 98, 1806 doi:10.1021/ja00423a027 ^ Baeckvall, J.E., Akermark, B., Ljunggren, S.O., J. Am. Chem. Soc., 1979, 101, 2411. doi:10.1021/ja00503a029 ^ Zaw, K., Lautens, M. and Henry P.M. Organometallics, 1985, 4, 1286-1296 ^ Wan W.K., Zaw K., and Henry P.M. Organometallics, 1988, 7, 1677-1683 ^ Francis, J.W., Henry, P.M. Organometallics, 1991, 10, 3498. doi:10.1021/om00056a019 ^ Francis, J.W., Henry, P.M. Organometallics, 1992, 11, 2832.doi:10.1021/om00044a024 ^ H. Stangl and R. Jira, Tetrahedron Lett., 1970, 11, 3589-3592 ^ T. Hosokawa, T. Nomura, S.-I. Murahashi, J. Organomet. Chem., 1998, 551, 387-389 ^ Comas-Vives, A., Stirling, A., Ujaque, G., Lleds, A., Chem. Eur. J., 2010, 16, 8738 8747.doi:10.1002/chem.200903522 ^ Anderson, B.J., Keith, J.A., and Sigman, M.S., J. Am. Chem. Soc., 2010, 132, 11872-11874 ^ J. A. Keith, J. Oxgaard, and W. A. Goddard, III J. Am. Chem. Soc., 2006, 128, 3132 - 3133; doi:10.1021/ja0533139 ^ H. E. Hosseini, S. A. Beyramabadi, A. Morsali, and M. R. Housaindokht, J. Mol. Struct. (THEOCHEM), 2010, 941, 138-143 ^ Jiro Tsuji, Hideo Nagashima, and Hisao Nemoto (1990), "General Synthetic Method for the preparation of Methyl Ketones from Terminal Olefins: 2-Decanone", Org. Synth., http://www.orgsyn.org/orgsyn/orgsyn/prepContent.asp?prep=cv7p0137; Coll. Vol. 7: 137 ^ McDonald, R.I., Liu, G., Stahl, S.S., J. Am. Chem. Soc., 2011, ASAPdoi:10.1021/cr100371y

27.

You might also like