Palladium(III) compounds

Last updated

In chemistry, compounds of palladium(III) feature the noble metal palladium in the unusual +3 oxidation state (in most of its compounds, palladium has the oxidation state II). Compounds of Pd(III) occur in mononuclear and dinuclear forms. Palladium(III) is most often invoked, not observed in mechanistic organometallic chemistry. [1] [2]

Contents

Mononuclear compounds

Pd(III) has a d7 electronic configuration, which leads to a Jahn–Teller distorted octahedral geometry. The geometry could also be viewed as being intermediate between square-planar and octahedral. These complexes are low-spin and paramagnetic.

Palladium(III) 1.png

The first Pd(III) complex characterized by X-ray crystallography was reported in 1987. [3] It was obtained by oxidation of the 1,4,7-trithiacyclononane (ttcn) complex [Pd(ttcn)2]3+. X-Ray crystallography revealed the expected Jahn–Teller distorted octahedral geometry, in spite of the highly symmetric structure of the ligand.

Palladium(III)2.png

The first organometallic Pd(III) complex characterized by X-Ray crystallography was reported in 2010. [4] Organopalladium complexes supported with a macrocyclic tetradentate ligand undergo single-electron oxidation to give Pd(III) species that is stabilized by the axially-positioned amine. The authors propose that while the axial nitrogen stabilize a distorted octahedral geometry, the t-Bu group and the rigidity of the macrocyclic structure inhibits the oxidation to a more conventional octahedral Pd(IV).

Palladium(III)3.png

Dinuclear compounds

Structure

Pairs of Pd(III) centers can couple, giving rise to a Pd–Pd bond. In contrast to the mononuclear Pd(III) complexes, the Pd(III)-Pd(III) dimers are diamagnetic.

Palladium(III)6.png

The first example of a dipalladium(III) complex was obtained by oxidation of dinuclear Pd(II) complex of triazabicyclodecene. [5]

Palladium(III)7.png

The first organometallic dinuclear Pd(III) complexes were reported in 2006 by Cotton and coworkers as well. [6] These complexes catalyze the diborylation of terminal olefins. [7] Due to the facile reduction of these complexes to Pd(II) species by diborane, the authors proposed that the dinuclear Pd(III) complexes serve as precatalysts for active Pd(II) catalysts.

Dipalladium(III) complexes derived from ortho-metallated triphenylphosphine (Ph groups omitted for clarity). Palladium(III)8.png
Dipalladium(III) complexes derived from ortho-metallated triphenylphosphine (Ph groups omitted for clarity).

Reactivity

The reactivity of dinuclear Pd(III) species as active catalytic intermediate is mostly discussed in the context of C-H activation. While it was proposed that Pd-catalyzed oxidative C-H functionalization reactions involve a Pd(IV) intermediate, Ritter and coworkers first postulated that these oxidative reactions could involve a dinuclear Pd(III) intermediate instead of Pd(IV). [8]

Palladium(III)9.png

Dinuclear Pd species are involved in Pd-catalyzed C-H chlorination. [9] Through X-ray crystallography, Ritter unambiguously showed that dinuclear Pd(III) complex is formed when the palladacycle is treated with two-electron oxidant, and such dinuclear complex undergoes C-Cl reductive elimination under ambient temperature. Both experimental and computational data was consistent with a concerted 1,1-reductive elimination mechanism for the C-Cl forming step. [10] [11] The authors show that such bimetallic participation of redox event lowers the activation barrier for reductive elimination step by ~30 kcal/mol compared to a monometallic pathway.

Palladium(III)10.png

Acetoxylation of 2-phenylpyridine was also demonstrated to involve a dinuclear Pd(III) intermediate. [12]

Related Research Articles

Reductive elimination is an elementary step in organometallic chemistry in which the oxidation state of the metal center decreases while forming a new covalent bond between two ligands. It is the microscopic reverse of oxidative addition, and is often the product-forming step in many catalytic processes. Since oxidative addition and reductive elimination are reverse reactions, the same mechanisms apply for both processes, and the product equilibrium depends on the thermodynamics of both directions.

<span class="mw-page-title-main">Wacker process</span>

The Wacker process or the Hoechst-Wacker process refers to the oxidation of ethylene to acetaldehyde in the presence of palladium(II) chloride and copper(II) chloride as the catalyst. This chemical reaction was one of the first homogeneous catalysis with organopalladium chemistry applied on an industrial scale.

<span class="mw-page-title-main">Pauson–Khand reaction</span> Chemical reaction

The Pauson–Khand (PK) reaction is a chemical reaction, described as a [2+2+1] cycloaddition. In it, an alkyne, an alkene and carbon monoxide combine into a α,β-cyclopentenone in the presence of a metal-carbonyl catalyst.

The Negishi coupling is a widely employed transition metal catalyzed cross-coupling reaction. The reaction couples organic halides or triflates with organozinc compounds, forming carbon-carbon bonds (C-C) in the process. A palladium (0) species is generally utilized as the metal catalyst, though nickel is sometimes used. A variety of nickel catalysts in either Ni0 or NiII oxidation state can be employed in Negishi cross couplings such as Ni(PPh3)4, Ni(acac)2, Ni(COD)2 etc.

<span class="mw-page-title-main">Transition metal dinitrogen complex</span> Coordination compounds with N2

Transition metal dinitrogen complexes are coordination compounds that contain transition metals as ion centers the dinitrogen molecules (N2) as ligands.

<span class="mw-page-title-main">Liebeskind–Srogl coupling</span>

The Liebeskind–Srogl coupling reaction is an organic reaction forming a new carbon–carbon bond from a thioester and a boronic acid using a metal catalyst. It is a cross-coupling reaction. This reaction was invented by and named after Jiri Srogl from the Academy of Sciences, Czech Republic, and Lanny S. Liebeskind from Emory University, Atlanta, Georgia, USA. There are three generations of this reaction, with the first generation shown below. The original transformation used catalytic Pd(0), TFP = tris(2-furyl)phosphine as an additional ligand and stoichiometric CuTC = copper(I) thiophene-2-carboxylate as a co-metal catalyst. The overall reaction scheme is shown below.

A transition metal oxo complex is a coordination complex containing an oxo ligand. Formally O2-, an oxo ligand can be bound to one or more metal centers, i.e. it can exist as a terminal or (most commonly) as bridging ligands (Fig. 1). Oxo ligands stabilize high oxidation states of a metal. They are also found in several metalloproteins, for example in molybdenum cofactors and in many iron-containing enzymes. One of the earliest synthetic compounds to incorporate an oxo ligand is potassium ferrate (K2FeO4), which was likely prepared by Georg E. Stahl in 1702.

Metal carbon dioxide complexes are coordination complexes that contain carbon dioxide ligands. Aside from the fundamental interest in the coordination chemistry of simple molecules, studies in this field are motivated by the possibility that transition metals might catalyze useful transformations of CO2. This research is relevant both to organic synthesis and to the production of "solar fuels" that would avoid the use of petroleum-based fuels.

A metal carbido complex is a coordination complex that contains a carbon atom as a ligand. They are analogous to metal nitrido complexes. Carbido complexes are a molecular subclass of carbides, which are prevalent in organometallic and inorganic chemistry. Carbido complexes represent models for intermediates in Fischer–Tropsch synthesis, olefin metathesis, and related catalytic industrial processes. Ruthenium-based carbido complexes are by far the most synthesized and characterized to date. Although, complexes containing chromium, gold, iron, nickel, molybdenum, osmium, rhenium, and tungsten cores are also known. Mixed-metal carbides are also known.

<span class="mw-page-title-main">Catellani reaction</span>

The Catellani reaction was discovered by Marta Catellani and co-workers in 1997. The reaction uses aryl iodides to perform bi- or tri-functionalization, including C-H functionalization of the unsubstituted ortho position(s), followed a terminating cross-coupling reaction at the ipso position. This cross-coupling cascade reaction depends on the ortho-directing transient mediator, norbornene.

In organic chemistry, the Fujiwara–Moritani reaction is a type of cross coupling reaction where an aromatic C-H bond is directly coupled to an olefinic C-H bond, generating a new C-C bond. This reaction is performed in the presence of a transition metal, typically palladium. The reaction was discovered by Yuzo Fujiwara and Ichiro Moritani in 1967. An external oxidant is required to this reaction to be run catalytically. Thus, this reaction can be classified as a C-H activation reaction, an oxidative Heck reaction, and a C-H olefination. Surprisingly, the Fujiwara–Moritani reaction was discovered before the Heck reaction.

Dialkylbiaryl phosphine ligands are phosphine ligands that are used in homogeneous catalysis. They have proved useful in Buchwald-Hartwig amination and etherification reactions as well as Negishi cross-coupling, Suzuki-Miyaura cross-coupling, and related reactions. In addition to these Pd-based processes, their use has also been extended to transformations catalyzed by nickel, gold, silver, copper, rhodium, and ruthenium, among other transition metals.

<span class="mw-page-title-main">Activation of cyclopropanes by transition metals</span>

In organometallic chemistry, the activation of cyclopropanes by transition metals is a research theme with implications for organic synthesis and homogeneous catalysis. Being highly strained, cyclopropanes are prone to oxidative addition to transition metal complexes. The resulting metallacycles are susceptible to a variety of reactions. These reactions are rare examples of C-C bond activation. The rarity of C-C activation processes has been attributed to Steric effects that protect C-C bonds. Furthermore, the directionality of C-C bonds as compared to C-H bonds makes orbital interaction with transition metals less favorable. Thermodynamically, C-C bond activation is more favored than C-H bond activation as the strength of a typical C-C bond is around 90 kcal per mole while the strength of a typical unactivated C-H bond is around 104 kcal per mole.

William B. Tolman an American inorganic chemist focusing on the synthesis and characterization of model bioinorganic systems, and organometallic approaches towards polymer chemistry. He has served as Editor in Chief of the ACS journal Inorganic Chemistry, and as a Senior Investigator at the NSF Center for Sustainable Polymers. Tolman is a Fellow of the American Association for the Advancement of Science and the American Chemical Society.

Palladacycle, as a class of metallacycles, refers to complexes containing at least one carbon-palladium bond. Palladacycles are invoked as intermediates in catalytic or palladium mediated reactions. They have been investigated as pre-catalysts for homogeneous catalysis and synthesis.

<span class="mw-page-title-main">Mizoroki-Heck vs. Reductive Heck</span>

The Mizoroki−Heck coupling of aryl halides and alkenes to form C(sp2)–C(sp2) bonds has become a staple transformation in organic synthesis, owing to its broad functional group compatibility and varied scope. In stark contrast, the palladium-catalyzed reductive Heck reaction has received considerably less attention, despite the fact that early reports of this reaction date back almost half a century. From the perspective of retrosynthetic logic, this transformation is highly enabling because it can forge alkyl–aryl linkages from widely available alkenes, rather than from the less accessible and/or more expensive alkyl halide or organometallic C(sp3) synthons that are needed in a classical aryl/alkyl cross-coupling.

β-Carbon elimination is a type of reaction in organometallic chemistry wherein an allyl ligand bonded to a metal center is broken into the corresponding metal-bonded alkyl (aryl) ligand and an alkene. It is a subgroup of elimination reactions. Though less common and less understood than β-hydride elimination, it is an important step involved in some olefin polymerization processes and transition-metal-catalyzed organic reactions.

Concerted metalation-deprotonation (CMD) is a mechanistic pathway through which transition-metal catalyzed C–H activation reactions can take place. In a CMD pathway, the C–H bond of the substrate is cleaved and the new C–Metal bond forms through a single transition state. This process does not go through a metal species that is bound to the cleaved hydrogen atom. Instead, a carboxylate or carbonate base deprotonates the substrate. The first proposal of a concerted metalation deprotonation pathway was by S. Winstein and T. G. Traylor in 1955 for the acetolysis of diphenylmercury. It was found to be the lowest energy transition state in a number of computational studies, was experimentally confirmed through NMR experiments, and has been hypothesized to occur in mechanistic studies.

Mono-N-protected amino acid (MPAA) is a bifunctional ligand that plays a key role in C–H functionalizations by accelerating the reaction rate and imparting specified chirality into the product. Amino acids are ideal building blocks for chiral ligand synthesis due to the cost, accessibility, large variety, solubility, and inherent chirality. Naturally occurring amino acids are transformed into chiral MPAA ligands that, upon coordination to metal complexes, allow reactions to occur that are otherwise energetically unfavorable. Great strides in the development of MPAA ligands over the past two decades have led to the integral role that enantioselective catalysis now plays in complex organic synthesis.

<span class="mw-page-title-main">Jirō Tsuji</span> Japanese chemist (1927–2022)

Jiro Tsuji was a Japanese chemist, notable for his discovery of organometallic reactions, including the Tsuji-Trost reaction, the Tsuji-Wilkinson decarbonylation, and the Tsuji-Wacker reaction.

References

  1. Mirica, Liviu M.; Khusnutdinova, Julia R. (2013-01-15). "Structure and electronic properties of Pd(III) complexes". Coordination Chemistry Reviews. 257 (2): 299–314. doi:10.1016/j.ccr.2012.04.030.
  2. Powers, David C.; Ritter, Tobias (2011). "Palladium(III) in Synthesis and Catalysis". In Canty, Allan J. (ed.). Higher Oxidation State Organopalladium and Platinum Chemistry. Topics in Organometallic Chemistry. Vol. 503. Springer Berlin Heidelberg. pp. 129–156. doi:10.1007/978-3-642-17429-2_6. ISBN   9783642174285. PMC   3066514 . PMID   21461129.
  3. Blake, Alexander J.; Holder, Alan J.; Hyde, Timothy I.; Schröder, Martin (January 1987). "Stabilisation of mononuclear palladium(III). The single crystal X-ray structure of the [Pd(L)2]3+cation (L = 1,4,7-trithiacyclononane)". J. Chem. Soc., Chem. Commun. (13): 987–988. doi:10.1039/c39870000987. ISSN   0022-4936.
  4. Khusnutdinova, Julia R.; Rath, Nigam P.; Mirica, Liviu M. (2010). "Stable Mononuclear Organometallic Pd(III) Complexes and Their C−C Bond Formation Reactivity". Journal of the American Chemical Society. 132 (21): 7303–7305. doi:10.1021/ja103001g. ISSN   0002-7863. PMID   20462195.
  5. Cotton, F. Albert; Gu, Jiande; Murillo, Carlos A.; Timmons, Daren J. (1998). "The First Dinuclear Complex of Palladium(III)". Journal of the American Chemical Society. 120 (50): 13280–13281. doi:10.1021/ja9832313. ISSN   0002-7863.
  6. Cotton, F. Albert; Koshevoy, Igor O.; Lahuerta, Pascual; Murillo, Carlos A.; Sanaú, Mercedes; Ubeda, M. Angeles; Zhao, Qinliang (2006). "High Yield Syntheses of Stable, Singly Bonded Pd26+ Compounds". Journal of the American Chemical Society. 128 (42): 13674–13675. doi:10.1021/ja0656595. ISSN   0002-7863. PMID   17044680.
  7. Penno, Dirk; Lillo, Vanesa; Koshevoy, Igor O.; Sanaú, Mercedes; Ubeda, M. Angeles; Lahuerta, Pascual; Fernández, Elena (2008). "Multifaceted Palladium Catalysts Towards the Tandem Diboration–Arylation Reactions of Alkenes". Chemistry – A European Journal. 14 (34): 10648–10655. doi:10.1002/chem.200800931. ISSN   1521-3765. PMID   18932177.
  8. Powers, David C.; Ritter, Tobias (2012). "Bimetallic Redox Synergy in Oxidative Palladium Catalysis". Accounts of Chemical Research. 45 (6): 840–850. doi:10.1021/ar2001974. ISSN   0001-4842. PMID   22029861.
  9. Powers, David C.; Ritter, Tobias (July 2009). "Bimetallic Pd(III) complexes in palladium-catalysed carbon–heteroatom bond formation". Nature Chemistry. 1 (4): 302–309. Bibcode:2009NatCh...1..302P. doi:10.1038/nchem.246. ISSN   1755-4330. PMID   21500602.
  10. Powers, David C.; Benitez, Diego; Tkatchouk, Ekaterina; Goddard, William A.; Ritter, Tobias (2010). "Bimetallic Reductive Elimination from Dinuclear Pd(III) Complexes". Journal of the American Chemical Society. 132 (40): 14092–14103. doi:10.1021/ja1036644. ISSN   0002-7863. PMC   2954233 . PMID   20858006.
  11. Powers, David C.; Xiao, Daphne Y.; Geibel, Matthias A. L.; Ritter, Tobias (2010). "On the Mechanism of Palladium-Catalyzed Aromatic C−H Oxidation". Journal of the American Chemical Society. 132 (41): 14530–14536. doi:10.1021/ja1054274. ISSN   0002-7863. PMC   2958923 . PMID   20873835.
  12. Powers, David C.; Geibel, Matthias A. L.; Klein, Johannes E. M. N.; Ritter, Tobias (2009). "Bimetallic Palladium Catalysis: Direct Observation of Pd(III)−Pd(III) Intermediates". Journal of the American Chemical Society. 131 (47): 17050–17051. doi:10.1021/ja906935c. ISSN   0002-7863. PMID   19899740. S2CID   207147534.