Palladium(II) acetate

Last updated
Palladium(II) acetate
Palladium(II)-acetate-2D.png
Pd(OAc)2.jpg
Pd(OAc)2-trimer-from-xtal-Mercury-3D-balls-A.png
Polymeric-Pd(OAc)2-from-xtal-2004-Mercury-3D-balls-A.png
Names
IUPAC name
Palladium(II) acetate
Other names
Palladium diacetate
hexakis(acetato)tripalladium
bis(acetato)palladium
Identifiers
3D model (JSmol)
ChemSpider
ECHA InfoCard 100.020.151 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 222-164-4
PubChem CID
RTECS number
  • AJ1900000
UNII
  • InChI=1S/2C2H4O2.Pd/c2*1-2(3)4;/h2*1H3,(H,3,4);/q;;+2/p-2 Yes check.svgY
    Key: YJVFFLUZDVXJQI-UHFFFAOYSA-L Yes check.svgY
  • InChI=1/2C2H4O2.Pd/c2*1-2(3)4;/h2*1H3,(H,3,4);/q;;+2/p-2
    Key: YJVFFLUZDVXJQI-NUQVWONBAH
  • ionic form:[Pd+2].[O-]C(=O)C.[O-]C(=O)C
  • coordination form (cyclic trimer):O0[C-](C)O[Pd+2]3(O[C-](C)O1)O[C-](C)O[Pd+2]1(O[C-](C)O2)O[C-](C)O[Pd-2]02O[C-](C)O3
Properties
Pd(CH3COO)2
Molar mass 224.51 g/mol
AppearanceBrown yellow solid
Density 2.19 g/cm3
Melting point 205 °C (401 °F; 478 K) decomposes
low
Structure
monoclinic
square planar
0 D
Hazards
Occupational safety and health (OHS/OSH):
Main hazards
considered nonhazardous
GHS labelling: [1]
GHS-pictogram-acid.svg GHS-pictogram-exclam.svg GHS-pictogram-pollu.svg
Danger
H317, H318, H410
P261, P272, P273, P280, P302+P352, P305+P351+P338
Safety data sheet (SDS)
Related compounds
Other anions
Palladium(II) chloride
Other cations
Platinum(II) acetate
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
Yes check.svgY  verify  (what is  Yes check.svgYX mark.svgN ?)

Palladium(II) acetate is a chemical compound of palladium described by the formula [Pd(O2CCH3)2]n, abbreviated [Pd(OAc)2]n. It is more reactive than the analogous platinum compound. Depending on the value of n, the compound is soluble in many organic solvents and is commonly used as a catalyst for organic reactions. [2]

Contents

Structure

With a 1:2 stoichiometric ratio of palladium atoms and acetate ligands, the compound exists as molecular and polymeric forms with the trimeric form being the dominant form in the solid state and in solution. Pd achieves approximate square planar coordination in both forms.

As prepared by Geoffrey Wilkinson and coworkers in 1965 and later characterized by Skapski and Smart in 1970 by single crystal X-ray diffraction, palladium(II) acetate is a red-brown solid that crystallizes as monoclinic plates. It has a trimeric structure, consisting of an equilateral triangle of Pd atoms each pair of which is bridged with two acetate groups in a butterfly conformation. [3] [4]

Palladium(II) acetate can also be prepared as a pale pink form. According to X-ray powder diffraction, this form is polymeric. [5]

Preparation

Palladium acetate, in trimeric form, can be prepared by treating palladium sponge with a mixture of acetic acid and nitric acid. An excess of palladium sponge metal or nitrogen gas flow are required to prevent contamination by the mixed nitrito-acetate (Pd3(OAc)5NO2). [6] [7]

Pd + 4 HNO3 → Pd(NO3)2 + 2 NO2 + 2 H2O
Pd(NO3)2 + 2 CH3COOH → Pd(O2CCH3)2 + 2 HNO3

Relative to the trimeric acetate, the mixed nitrate-acetate variant has different solubility and catalytic activity. Preventing, or controlling for the amount of, this impurity can be an important aspect for reliable use of palladium(II) acetate. [8]

Palladium(II) propionate is prepared analogously; other carboxylates are prepared by treating palladium(II) acetate with the appropriate carboxylic acid. [3] Likewise, palladium(II) acetate can be prepared by treating other palladium(II) carboxylates with acetic acid. This ligand exchange starting with a purified other carboxylate is an alternative way to synthesize palladium(II) acetate free from the nitro contaminant. [8]

Palladium(II) acetate is prone to reduction to Pd(0) in the presence of reagents which can undergo beta-hydride elimination such as primary and secondary alcohols as well as amines. When warmed with alcohols, or on prolonged boiling with other solvents, palladium(II) acetate decomposes to palladium. [3]

Catalysis

Palladium acetate is a catalyst for many organic reactions, especially alkenes, dienes, and alkyl, aryl, and vinyl halides to form reactive adducts. [9]

Reactions catalyzed by palladium(II) acetate:

RC6H4Br + Si2(CH3)6 → RC6H4Si(CH3)3 + Si(CH3)3Br

Pd(O2CCH3)2 is compatible with the electronic properties of aryl bromides, and unlike other methods of synthesis, this method does not require high pressure equipment. [15]

Precursor to other Pd compounds

Palladium acetate is used to produce other palladium(II) compounds. For example, phenylpalladium acetate, used to isomerize allyl alcohols to aldehydes, is prepared by the following reaction: [16]

Hg(C6H5)(OAc) + Pd(OAc)2 → Pd(C6H5)(OAc) + Hg(OAc)2

Palladium(II) acetate reacts with acetylacetone (the "acac" ligand) to produce Pd(acac)2.

Herrmann's catalyst is made by reaction of palladium(II) acetate with tris(o-tolyl)phosphine. [17]

Structure of Herrmann's catalyst. HerrmannCat.png
Structure of Herrmann's catalyst.

Light or heat reduce palladium acetate to give thin layers of palladium and can produce nanowires and colloids. [6]

See also

Related Research Articles

<span class="mw-page-title-main">Ether</span> Organic compounds made of alkyl/aryl groups bound to oxygen (R–O–R)

In organic chemistry, ethers are a class of compounds that contain an ether group—an oxygen atom connected to two organyl groups. They have the general formula R−O−R′, where R and R′ represent organyl groups. Ethers can again be classified into two varieties: if the organyl groups are the same on both sides of the oxygen atom, then it is a simple or symmetrical ether, whereas if they are different, the ethers are called mixed or unsymmetrical ethers. A typical example of the first group is the solvent and anaesthetic diethyl ether, commonly referred to simply as "ether". Ethers are common in organic chemistry and even more prevalent in biochemistry, as they are common linkages in carbohydrates and lignin.

<span class="mw-page-title-main">Ester</span> Compound derived from an acid

In chemistry, an ester is a compound derived from an acid in which the hydrogen atom (H) of at least one acidic hydroxyl group of that acid is replaced by an organyl group. Analogues derived from oxygen replaced by other chalcogens belong to the ester category as well. According to some authors, organyl derivatives of acidic hydrogen of other acids are esters as well, but not according to the IUPAC.

The Heck reaction is the chemical reaction of an unsaturated halide with an alkene in the presence of a base and a palladium catalyst to form a substituted alkene. It is named after Tsutomu Mizoroki and Richard F. Heck. Heck was awarded the 2010 Nobel Prize in Chemistry, which he shared with Ei-ichi Negishi and Akira Suzuki, for the discovery and development of this reaction. This reaction was the first example of a carbon-carbon bond-forming reaction that followed a Pd(0)/Pd(II) catalytic cycle, the same catalytic cycle that is seen in other Pd(0)-catalyzed cross-coupling reactions. The Heck reaction is a way to substitute alkenes.

The Sonogashira reaction is a cross-coupling reaction used in organic synthesis to form carbon–carbon bonds. It employs a palladium catalyst as well as copper co-catalyst to form a carbon–carbon bond between a terminal alkyne and an aryl or vinyl halide.

Organopalladium chemistry is a branch of organometallic chemistry that deals with organic palladium compounds and their reactions. Palladium is often used as a catalyst in the reduction of alkenes and alkynes with hydrogen. This process involves the formation of a palladium-carbon covalent bond. Palladium is also prominent in carbon-carbon coupling reactions, as demonstrated in tandem reactions.

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

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">Tetrakis(triphenylphosphine)palladium(0)</span> Chemical compound

Tetrakis(triphenylphosphine)palladium(0) (sometimes called quatrotriphenylphosphine palladium) is the chemical compound [Pd(P(C6H5)3)4], often abbreviated Pd(PPh3)4, or rarely PdP4. It is a bright yellow crystalline solid that becomes brown upon decomposition in air.

<span class="mw-page-title-main">Organomercury chemistry</span> Group of chemical compounds containing mercury

Organomercury chemistry refers to the study of organometallic compounds that contain mercury. Typically the Hg–C bond is stable toward air and moisture but sensitive to light. Important organomercury compounds are the methylmercury(II) cation, CH3Hg+; ethylmercury(II) cation, C2H5Hg+; dimethylmercury, (CH3)2Hg, diethylmercury and merbromin ("Mercurochrome"). Thiomersal is used as a preservative for vaccines and intravenous drugs.

In organic chemistry, the Buchwald–Hartwig amination is a chemical reaction for the synthesis of carbon–nitrogen bonds via the palladium-catalyzed coupling reactions of amines with aryl halides. Although Pd-catalyzed C–N couplings were reported as early as 1983, Stephen L. Buchwald and John F. Hartwig have been credited, whose publications between 1994 and the late 2000s established the scope of the transformation. The reaction's synthetic utility stems primarily from the shortcomings of typical methods for the synthesis of aromatic C−N bonds, with most methods suffering from limited substrate scope and functional group tolerance. The development of the Buchwald–Hartwig reaction allowed for the facile synthesis of aryl amines, replacing to an extent harsher methods while significantly expanding the repertoire of possible C−N bond formations.

<span class="mw-page-title-main">1,1'-Bis(diphenylphosphino)ferrocene</span> Chemical compound

1,1-Bis(diphenylphosphino)ferrocene, commonly abbreviated dppf, is an organophosphorus compound commonly used as a ligand in homogeneous catalysis. It contains a ferrocene moiety in its backbone, and is related to other bridged diphosphines such as 1,2-bis(diphenylphosphino)ethane (dppe).

In chemistry, carbonylation refers to reactions that introduce carbon monoxide (CO) into organic and inorganic substrates. Carbon monoxide is abundantly available and conveniently reactive, so it is widely used as a reactant in industrial chemistry. The term carbonylation also refers to oxidation of protein side chains.

The Fukuyama coupling is a coupling reaction taking place between a thioester and an organozinc halide in the presence of a palladium catalyst. The reaction product is a ketone. This reaction was discovered by Tohru Fukuyama et al. in 1998.

<span class="mw-page-title-main">Organobismuth chemistry</span>

Organobismuth chemistry is the chemistry of organometallic compounds containing a carbon to bismuth chemical bond. Applications are few. The main bismuth oxidation states are Bi(III) and Bi(V) as in all higher group 15 elements. The energy of a bond to carbon in this group decreases in the order P > As > Sb > Bi. The first reported use of bismuth in organic chemistry was in oxidation of alcohols by Frederick Challenger in 1934 (using Ph3Bi(OH)2). Knowledge about methylated species of bismuth in environmental and biological media is limited.

<span class="mw-page-title-main">Bis(triphenylphosphine)palladium chloride</span> Chemical compound

Bis(triphenylphosphine)palladium chloride is a coordination compound of palladium containing two triphenylphosphine and two chloride ligands. It is a yellow solid that is soluble in some organic solvents. It is used for palladium-catalyzed coupling reactions, e.g. the Sonogashira–Hagihara reaction. The complex is square planar. Many analogous complexes are known with different phosphine ligands.

<span class="mw-page-title-main">(1,1'-Bis(diphenylphosphino)ferrocene)palladium(II) dichloride</span> Chemical compound

[1,1'‑Bis(diphenylphosphino)ferrocene]palladium(II) dichloride is a palladium complex containing the bidentate ligand 1,1'-bis(diphenylphosphino)ferrocene (dppf), abbreviated as [(dppf)PdCl2]. This commercially available material can be prepared by reacting dppf with a suitable nitrile complex of palladium dichloride:

<span class="mw-page-title-main">White catalyst</span> Chemical compound

The White catalyst is a transition metal coordination complex named after the chemist by whom it was first synthesized, M. Christina White, a professor at the University of Illinois. The catalyst has been used in a variety of allylic C-H functionalization reactions of α-olefins. In addition, it has been shown to catalyze oxidative Heck reactions.

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">Herrmann's catalyst</span> Organopalladium compound used as a catalyst

Herrmann's catalyst is an organopalladium compound that is a popular catalyst for the Heck reaction. It is a yellow air-stable solid that is soluble in organic solvents. Under conditions for catalysis, the acetate group is lost and the Pd-C bond undergoes protonolysis, giving rise to a source of "PdP(o-tol)3".

<span class="mw-page-title-main">Transition metal carboxylate complex</span> Class of chemical compounds

Transition metal carboxylate complexes are coordination complexes with carboxylate (RCO2) ligands. Reflecting the diversity of carboxylic acids, the inventory of metal carboxylates is large. Many are useful commercially, and many have attracted intense scholarly scrutiny. Carboxylates exhibit a variety of coordination modes, most common are κ1- (O-monodentate), κ2 (O,O-bidentate), and bridging.

Palladium forms a variety of ionic, coordination, and organopalladium compounds, typically with oxidation state Pd0 or Pd2+. Palladium(III) compounds have also been reported. Palladium compounds are frequently used as catalysts in cross-coupling reactions such as the Sonogashira coupling and Suzuki reaction.

References

  1. "520764 [Pd(OAc)2]3". Sigma-Aldrich. Retrieved 23 December 2021.
  2. Grennberg, Helena; Foot, Jonathan S.; Banwell, Martin G.; Roman, Daniela Sustac (2001). "Palladium(II) Acetate". Encyclopedia of Reagents for Organic Synthesis. pp. 1–35. doi:10.1002/047084289X.rp001.pub3. ISBN   978-0-470-84289-8.
  3. 1 2 3 T. A. Stephenson; S. M. Morehouse; A. R. Powell; J. P. Heffer; G. Wilkinson (1965). "667. Carboxylates of palladium, platinum, and rhodium, and their adducts". Journal of the Chemical Society (Resumed): 3632. doi:10.1039/jr9650003632.
  4. Skapski, A C.; M. L. Smart (1970). "The Crystal Structure of Trimeric Palladium(II) Acetate". J. Chem. Soc. D (11): 658b–659. doi:10.1039/C2970000658b.
  5. Kirik, S.D.; Mulagaleev, S.F.; Blokhin, A.I. (2004). "[Pd(CH3COO)2]n from X-ray Powder Diffraction Data". Acta Crystallogr. C . 60 (9): m449–m450. doi:10.1107/S0108270104016129. PMID   15345831.
  6. 1 2 Bakhmutov, V. I.; Berry, J. F.; Cotton, F. A.; Ibragimov, S.; Murillo, C. A. (2005). "Non-Trivial Behavior of Palladium(II) Acetate". Dalton Transactions (11): 1989–1992. doi:10.1039/b502122g. PMID   15909048.
  7. "High Purity Homogeneous Catalyst" (PDF). Engelhard. September 2005. Archived from the original (PDF) on 17 March 2006. Retrieved 24 February 2006.
  8. 1 2 Ritter, Stephen K. (May 2, 2016). "Chemists introduce a user's guide for palladium acetate". Chemical & Engineering News . 94 (18): 20–21. doi:10.1021/cen-09418-scitech1.
  9. Suggs, J W. "Palladium: Organometallic Chemistry." Encyclopedia of Inorganic Chemistry. Ed. R B. King. 8 vols. Chichester: Wiley, 1994.
  10. Keary M. Engle; Navid Dastbaravardeh; Peter S. Thuy-Boun; Dong-Hui Wang; Aaron C. Sather; Jin-Quan Yu (2015). "Ligand-Accelerated ortho-C-H Olefination of Phenylacetic Acids". Org. Synth. 92: 58–75. doi:10.15227/orgsyn.092.0058. PMC   4936495 . PMID   27397943.
  11. Nikitin, Kirill V.; Andryukhova, N.P.; Bumagin, N.A.; Beletskaya, I.P. (1991). "Synthesis of Aryl Esters by Pd-catalysed Carbonylation of Aryl Iodides". Mendeleev Communications. 1 (4): 129–131. doi:10.1070/MC1991v001n04ABEH000080.
  12. Basu, B., Satadru J., Mosharef H. B., and Pralay D. (2003). "A Simple Protocol for the Direct Reductive Amination of Aldehydes and Ketones Using Potassium Formate and Catalytic Palladium Acetate". ChemInform . 34 (30): 555–557. doi:10.1002/chin.200330069.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  13. Linli He; Shawn P. Allwein; Benjamin J. Dugan; Kyle W. Knouse; Gregory R. Ott; Craig A. Zificsak (2016). "Synthesis of α-Carboline". Org. Synth. 93: 272. doi: 10.15227/orgsyn.093.0272 .
  14. "Buchwald-Hartwig Cross Coupling Reaction". Organic Chemistry Portal.
  15. Gooben, L J. "Research Area "New Pd-Catalyzed Cross-Coupling Reactions"" 28 Feb. 2006<http://www.mpi-muelheim.mpg.de/kofo/bericht2002/pdf/2.1.8_gossen.pdf> Archived July 12, 2007, at the Wayback Machine
  16. Richard F. Heck. "Aldehydes from Allylic Alcohols and Phenylpalladium Acetate: 2-Methyl-3-Phenylpropionaldehyde". Organic Syntheses ; Collected Volumes, vol. 6, p. 815.
  17. Herrmann, W. A.; Brossmer, C.; Reisinger, C.-P.; Riermeier, T. H.; Öfele, K.; Beller, M. (1997). "Palladacycles: Efficient New Catalysts for the Heck Vinylation of Aryl Halides". Chemistry – A European Journal. 3: 1357–1364. doi:10.1002/chem.19970030823.
  18. Herrmann, W. A.; Brossmer, C.; Reisinger, C.-P.; Riermeier, T. H.; Öfele, K.; Beller, M. (1997). "Palladacycles: Efficient New Catalysts for the Heck Vinylation of Aryl Halides". Chemistry – A European Journal. 3 (8): 1357–1364. doi:10.1002/chem.19970030823.