Fact-checked by Grok 2 weeks ago

Wacker process

The Wacker process is an industrial catalytic oxidation reaction that converts ethylene (ethene) into acetaldehyde using a palladium(II) chloride and copper(II) chloride catalyst system in the presence of water and molecular oxygen as the terminal oxidant. Developed in the late 1950s by chemists at Wacker Chemie and Hoechst AG, the process marked a significant advancement in homogeneous catalysis, enabling efficient production of acetaldehyde—a key intermediate in the synthesis of acetic acid, acetic anhydride, and other chemicals—from inexpensive petroleum-derived ethylene. The reaction proceeds under mild conditions, typically at 50–130°C and 1–10 atm pressure in an containing to maintain catalyst stability, with the process operating in a two-stage manner: the initial palladium-catalyzed step forms and reduces Pd(II) to Pd(0), followed by reoxidation of Pd(0) via Cu(II) and regeneration of Cu(II) by oxygen. The involves coordination of to Pd(II), syn addition of water to form a β-hydroxyalkylpalladium , and subsequent β-hydride elimination to yield , ensuring high selectivity (>95%) and yields up to 98% under optimized conditions. This avoids the need for stoichiometric oxidants, making the process economically viable despite the relatively high cost of . Industrially, the Wacker process revolutionized production, replacing older methods like and contributing to the shift from coal-based to petroleum-based feedstocks in the ; initial plants in had capacities of around 15,000 tons per year, rapidly expanding thereafter with operational costs competitive against alternatives. Today, it remains a cornerstone for producing over 1 million tons of annually worldwide (as of 2024), primarily for downstream applications in , solvents, and polymers, though its scope has expanded to lab-scale variants like the Tsuji-Wacker oxidation for converting higher alkenes to ketones. Ongoing research focuses on sustainable improvements, such as direct oxygen coupling without copper or heterogeneous catalysts, to enhance efficiency and reduce environmental impact.

Historical Background

Discovery

The discovery of the Wacker process originated from research conducted at in during the early to mid-1950s, as part of efforts to develop efficient methods for oxidizing to using readily available feedstocks. A team led by Julius Smidt, including Reinhard Jira, Walter Hafner, and others, initiated experiments exploring palladium-based catalysis for this transformation. Initial studies demonstrated that reacts with PdCl₂ in to yield , but this was limited to a stoichiometric process where palladium was consumed and required regeneration. Key progress came from identifying copper(II) chloride (CuCl₂) as an essential co-catalyst, which facilitates the reoxidation of reduced palladium species using molecular oxygen (O₂), enabling a catalytic cycle. This innovation overcame early challenges, such as palladium deactivation and inefficient oxygen utilization, by stabilizing the system in an acidic aqueous medium and preventing copper precipitation through careful control of chloride concentrations and pH. The complete catalytic system—PdCl₂/CuCl₂ in water under oxygen atmosphere—achieved high selectivity for acetaldehyde, marking a breakthrough in homogeneous catalysis. The feasibility of the process was first detailed in a seminal 1959 publication by Smidt, Hafner, , Sedlmeier, Sieber, Rüttinger, and Kojer in , which outlined the reaction conditions and preliminary yields exceeding 90% based on conversion. Supporting patents, including applications filed starting in 1956, protected the invention and demonstrated its viability through lab-scale experiments. These developments laid the foundation for industrial application, as recounted in a 2009 retrospective by , one of the original researchers.

Commercialization

The commercialization of the Wacker process marked a significant advancement in , transitioning from laboratory-scale experiments to large-scale shortly after its discovery. Wacker Chemie launched the first commercial plant in Cologne-Merkenich, Germany, in 1960, with an initial annual capacity of 15,000 metric tons of . This facility represented the initial application of the ethylene-based oxidation, replacing earlier acetylene-derived methods developed by the company. The rapid scaling was facilitated by key patents granted to , which protected the palladium-catalyzed process and enabled swift engineering adaptations for continuous operation. Licensing agreements played a crucial role in the global adoption of the technology during the . Wacker Chemie, in collaboration with Hoechst through the jointly owned , granted licenses to multiple international firms, including Corporation, which opened an plant in , in 1962 using the process. By 1960, the technology had been licensed 17 times worldwide, promoting widespread implementation in regions with access to low-cost feedstocks. Although specific details on DuPont's involvement are limited, the process's proliferation reflected strategic partnerships that accelerated market entry beyond . The economic drivers behind the Wacker process's success stemmed from its efficiency compared to prior acetaldehyde production routes, such as the oxidation of or hydration of . By directly utilizing abundant and inexpensive derived from cracking, the process achieved substantial cost savings in production expenses while improving and reducing . This shift not only displaced less selective ethanol oxidation methods, which suffered from lower conversion rates and higher byproduct formation, but also aligned with the post-World War II expansion of infrastructure, solidifying 's role as a key intermediate for , , and production.

Reaction Overview

General Description

The Wacker process is a catalytic aerobic oxidation reaction that converts ethylene into acetaldehyde, serving as a cornerstone of industrial organic synthesis for producing carbonyl compounds from alkenes. Developed in the mid-20th century, it employs palladium(II) chloride as the primary catalyst and copper(II) chloride as a co-catalyst, enabling efficient use of molecular oxygen as the terminal oxidant. This process selectively transforms terminal alkenes, such as ethylene, into the corresponding methyl ketones or aldehydes, with acetaldehyde being the prototypical product. The overall transformation is represented by the net equation: \ce{C2H4 + 1/2 O2 -> CH3CHO} This reaction occurs in an aqueous medium, where the catalysts facilitate the incorporation of oxygen into the without the need for stoichiometric oxidants. Stoichiometrically, the process begins with the oxidation of by PdCl₂ in the presence of , producing and reducing Pd(II) to Pd(0), followed by reoxidation of Pd(0) using CuCl₂ to regenerate PdCl₂ and reduce Cu(II) to Cu(I); the Cu(I) is then reoxidized to Cu(II) by molecular oxygen in acidic conditions. This cycle ensures catalytic turnover of both metals, achieving high selectivity and efficiency. Classified as an aerobic oxidation, the Wacker process exemplifies the use of transition metals to activate O₂ for selective C-H functionalization in alkene substrates, contrasting with non-catalytic direct oxidations of ethylene using pure oxygen, which predominantly yield ethylene oxide via epoxidation pathways.

Catalysts and Reaction Conditions

The primary catalysts in the Wacker process are palladium(II) chloride (PdCl₂) as the active species and copper(II) chloride (CuCl₂) as the co-catalyst and reoxidant, with PdCl₂ employed at catalytic concentrations relative to ethylene and CuCl₂ in excess to facilitate reoxidation. The reaction medium consists of an aqueous hydrochloric acid solution to provide the necessary acidic conditions. Reaction conditions typically involve temperatures of 50–130°C and pressures ranging from 1 to 10 , with continuously bubbled through the catalyst solution and molecular oxygen (or air in certain configurations) as the terminal oxidant to regenerate the copper species. Chloride ions from the HCl are essential for stabilizing soluble palladium complexes, such as [PdCl₄]²⁻, which prevents the precipitation of inactive palladium(0) species during the . Solvent variations are employed depending on the ; aqueous media suit due to its , while acetic acid is used for less soluble higher alkenes to enhance and favor ester formation as an . Pure O₂ is generally preferred over air to mitigate explosion hazards from flammable gas mixtures in the reactor.

Reaction Mechanism

Key Steps

The Wacker process begins with the coordinative addition of Pd(II), typically as PdCl₂, to , forming a π-complex denoted as (η²-C₂H₄)PdCl₂. This coordination activates the for subsequent nucleophilic attack. Following this, the activated undergoes nucleophilic attack by on one of the carbons, with the bonding to the other carbon, resulting in to form the trans-β-hydroxyalkylpalladium intermediate, trans-(HO-CH₂-CH₂)PdCl₂. The of the hydroxypalladation is under typical high- industrial conditions, though predominates at low chloride concentrations. This step establishes the carbon-oxygen bond essential for the eventual carbonyl formation. The hydroxy-palladated intermediate then undergoes β-hydride elimination, transferring a from the β-carbon to , producing the enol form of coordinated to Pd(0). The enol tautomerizes rapidly to (CH₃CHO) under the reaction conditions. To close the cycle, reoxidation is required: Pd(0) is oxidized back to Pd(II) by Cu(II) chloride, generating Cu(I). Subsequently, Cu(I) is reoxidized to Cu(II) by molecular oxygen (O₂), producing as a . This two-step reoxidation ensures catalytic turnover without net consumption of the metal catalysts. The overall catalytic cycle achieves the net transformation 2 C₂H₄ + O₂ → 2 CH₃CHO, enabling efficient aerobic oxidation.

Supporting Evidence

Kinetic studies conducted in the 1950s and 1960s demonstrated that the rate of the Wacker process exhibits dependence on both the concentration of Pd(II) and , while showing an inverse dependence on concentration, typically inverse second-order at low levels. These findings, derived from detailed rate law determinations, support the involvement of a Pd(II)-olefin as a key intermediate and highlight 's role in inhibiting the reaction through coordination to . Isotopic labeling experiments using H₂¹⁸O have confirmed that water serves as the source of the oxygen atom incorporated into the product, rather than molecular oxygen. In these studies, the labeled oxygen from water was detected in the of , providing direct evidence for the hydroxypalladation step where water acts as the attacking the Pd(II)-bound olefin. Spectroscopic investigations have offered further validation of the mechanistic intermediates. UV-Vis spectroscopy has identified characteristic absorption bands for Pd-olefin π-complexes, confirming their formation under reaction conditions and supporting the initial coordination step. Additionally, electron spin resonance (ESR) spectroscopy has detected Cu(II) species, such as CuCl₂, illustrating their role in reoxidizing Pd(0) and the overall catalytic cycle. Hammett studies on substituted styrenes have reinforced the electrophilic nature of the palladium addition to the olefin. These investigations revealed a positive ρ value (approximately +0.5 to +1.0), indicating that electron-donating substituents on the styrene accelerate the reaction, consistent with rate-determining nucleophilic attack by water on the electron-deficient . Modern computational studies using (DFT), conducted after 2000, have validated the energy profiles of critical steps including syn-addition of water to the Pd-olefin complex and subsequent β-hydride elimination. These calculations predict low barriers (around 20-30 kcal/mol) for the outer-sphere hydroxypalladation pathway under aqueous conditions, aligning with experimental kinetics and resolving earlier debates on versus anti addition mechanisms.

Industrial Processes

One-Stage Process

The one-stage process, also known as the Wacker-Hoechst process, for the industrial Wacker oxidation integrates the palladium-catalyzed oxidation of ethylene and the copper-mediated reoxidation in a single reactor, enabling continuous production of acetaldehyde. Ethylene, oxygen, and recycle streams from unreacted gases are fed into the lower part of a reaction tower containing an aqueous catalyst solution of palladium(II) chloride, copper(II) chloride, and hydrochloric acid. The gases bubble through the catalyst solution, where ethylene is oxidized to acetaldehyde, while the reduced palladium is reoxidized by copper(II), which is in turn regenerated by oxygen. The reaction mixture is circulated via an airlift principle to a separating vessel for disengagement, and the product-laden vapor is cooled, scrubbed with water to recover acetaldehyde, and sent to distillation, with unconverted gases recycled back to the reactor. The process operates at temperatures of 100–130 °C and pressures of 4–10 bar to achieve high reaction rates and solubility of gases in the aqueous medium. As detailed in the catalysts and reaction conditions section, the Pd/Cu/HCl system facilitates the overall aerobic oxidation. A portion of the recycle stream is continuously vented to purge inert gases and maintain steady-state operation. Ethylene conversion exceeds 95%, with acetaldehyde yields approaching the same level due to efficient recycling and minimal side reactions under optimized conditions. This setup offers advantages in simplified design and lower capital costs compared to multi-reactor configurations, as it avoids intermediate separations and allows for compact, continuous operation in a single vessel. The integrated approach also minimizes energy losses from gas handling between stages, contributing to economic viability in large-scale production. However, challenges include severe from the acidic HCl generated in the reaction, necessitating specialized materials such as , , or ceramic-lined reactors and piping. Additionally, the mixture of and oxygen poses explosion risks, which are mitigated by limiting the oxygen concentration in the feed to below 8% to stay outside the flammable limits. Byproducts are minimal, primarily consisting of from incomplete traces, along with small amounts of chlorinated hydrocarbons and formed via over-oxidation or side reactions. The is recovered by from the top of the column after scrubbing, achieving high purity (>99%) through extractive and steps to remove light ends and heavies.

Two-Stage Process

The two-stage process, also known as the Veba-Chemie process, employs sequential reactors to perform the palladium- and copper-catalyzed oxidation of ethylene to acetaldehyde, enabling better separation of reaction components and improved operational control compared to integrated approaches. In the first stage, ethylene is oxidized in a tubular reactor using an aqueous solution of palladium(II) chloride (PdCl₂) and copper(II) chloride (CuCl₂) as catalysts, typically at 90–100°C and 10 atm pressure, yielding acetaldehyde while reducing Cu(II) to Cu(I). The acetaldehyde is then separated via flashing and distillation, leaving behind the reduced catalyst mixture. In the second stage, the Cu(I)-containing solution is transferred to a separate vessel for reoxidation with air or oxygen to regenerate , which in turn reoxidizes Pd(0) back to ; this step often occurs in low-acid conditions to suppress side reactions like over-oxidation to acetic acid. The regenerated catalyst solution is recycled to the first reactor, while off-gases are scrubbed to recover trace . This staged design offers key advantages, including selectivity greater than 95% for , reduced equipment corrosion through segregated acidic and oxidative environments (necessitating lined reactors and piping only where essential), and simplified by of any insoluble before . The process also permits the use of air rather than pure oxygen, lowering costs and risks. Commercialized by in 1960, the two-stage variant was licensed to companies including and starting in the early , supporting plants with typical annual capacities exceeding 100,000 metric tons of . At its peak, global production via such processes reached over 900,000 metric tons per year.

Variants and Extensions

Tsuji-Wacker Oxidation

The Tsuji-Wacker oxidation represents an intramolecular adaptation of the Wacker process, pioneered by Tsuji during the and , wherein Pd(II) catalyzes the oxidation of internal alkenes bearing tethered nucleophiles to yield cyclic carbonyl compounds. This variant extends the utility of palladium-mediated oxidations beyond simple terminal alkenes, enabling efficient construction of ring systems through nucleophilic capture of palladated intermediates. The reaction's general scope encompasses the transformation of alkenols into lactones, alkenamines into lactams, and related substrates into cyclic ketones, proceeding under mild aerobic conditions that leverage molecular oxygen as the terminal oxidant. Typical protocols employ as the precatalyst, CuCl₂ as a co-oxidant, and O₂ in solvents such as DMF or aqueous media to facilitate reoxidation of Pd(0) to Pd(II). These conditions promote high functional group tolerance and scalability for synthetic applications, distinguishing the process from stoichiometric palladium oxidations. In contrast to the classical intermolecular Wacker process, the Tsuji variant's intramolecular design imparts inherent , favoring the formation of 5- or 6-membered rings via directed nucleophilic attack on the coordinated . A representative example is the conversion of 4-penten-1-ol to , wherein the pendant nucleophile cyclizes onto the activated to afford the 5-membered in good yield. The underlying mirrors that of the classical process, involving syn addition of Pd(II) and a nucleophilic species across the , followed by β-hydride elimination and catalyst turnover.

Modern Developments

Since the early , ligand-modified catalysts have enabled greater control over in Wacker-type oxidations, particularly achieving anti-Markovnikov addition for terminal alkenes to yield . For instance, co-catalysts with Pd(II) promote aldehyde formation from terminal alkenes by favoring external nucleophilic attack by water, with yields often exceeding 80% under mild conditions. This approach, developed in the late and refined through the 2010s, contrasts with traditional Markovnikov selectivity and has been applied to aliphatic and aromatic substrates, demonstrating high for aldehyde products. Efforts to enhance sustainability have led to copper-free Wacker systems, eliminating the need for toxic co-catalysts while maintaining efficiency. Copper-free variants using molecular oxygen or peroxides as oxidants have been reported, enabling aerobic conditions with reduced waste. Further advancements incorporate direct O₂ as the terminal oxidant with nitrogen-based ligands on Pd complexes, enabling aerobic conditions at and temperatures around 80°C, with selectivities above 90% for methyl ketones from terminal alkenes. These systems prioritize principles, minimizing metal contamination and enabling catalyst loadings as low as 0.1 mol%. Intermolecular Wacker oxidations have expanded in the to include enantioselective variants for styrene derivatives, producing aryl methyl ketones with enantiomeric excess. Chiral bidentate ligands enable asymmetric during nucleopalladation under O₂-mediated conditions. This development supports the of chiral building blocks for pharmaceuticals. Recent applications integrate Wacker oxidations into flow chemistry platforms and bio-compatible media, facilitating scalable from 2022 onward. Continuous-flow setups with Pd catalysts allow precise control of O₂ delivery, yielding high conversions for precursors while enhancing safety compared to batch methods. Bio-compatible variants employ aqueous buffers and H₂O₂ oxidants at neutral pH, enabling late-stage functionalization of complex scaffolds, as demonstrated in syntheses of oxygenated terpenoids. Emerging bio-inspired approaches, such as of P450 enzymes for anti-Markovnikov Wacker-type oxidations, offer sustainable alternatives as of 2024. Environmental enhancements focus on waste reduction and , with air or O₂ as oxidants and recyclable Pd nanoparticles. Supported Pd nanoparticles enable cocatalyst-free aerobic Wacker oxidations, producing as the sole . These systems align with industrial goals.

Applications and Impact

Industrial Uses

The Wacker process plays a central role in the industrial production of , serving as the primary method for converting into this key intermediate chemical. produced via this process is predominantly used as a feedstock for manufacturing acetic acid, , , and , which are essential for applications in polymers, adhesives, and solvents. Global acetaldehyde production capacity stands at approximately 1.19 million tons annually as of 2025, with projections for steady growth driven by demand in these downstream sectors. Despite its efficiency, the overall scale of acetaldehyde production—and thus the Wacker process—has declined relative to historical levels due to the widespread adoption of alternative routes for acetic acid synthesis, such as methanol carbonylation, which eliminates the need for acetaldehyde as an intermediate. Nevertheless, the Wacker process accounts for more than 85% of global acetaldehyde output and remains a cornerstone in regions like Asia and Europe, where petrochemical infrastructure supports its continued viability. Major industrial implementations of the Wacker process are operated by companies including Celanese Corporation in the United States and in , with facilities optimized for high-yield oxidation. Some plants have been adapted to utilize the process for higher alkenes, such as the oxidation of to acetone, enhancing versatility in bulk chemical manufacturing. The process's energy profile is favorable compared to legacy methods like acetylene hydration, contributing to its economic sustainability in modern complexes.

Synthetic Applications

The Wacker and Tsuji-Wacker oxidations have found extensive application in the of complex natural products, particularly for the regioselective conversion of allylic alcohols or terminal alkenes to carbonyl compounds under mild conditions. In Teruaki Mukaiyama's asymmetric of taxol, a Wacker-type oxidation was employed to transform an allylic alcohol intermediate into the corresponding diketone, facilitating the construction of the ABC ring system essential to the molecule's structure. Similarly, in efforts toward aglycons and related glycopeptides, Tsuji-Wacker oxidation has been utilized to generate key tricyclic intermediates from sulfamate-tethered precursors, enabling the formation of the rigid framework with high stereocontrol. These applications, spanning the 1990s to the 2020s, highlight the reaction's role in assembling polycyclic scaffolds in medicinally important targets. Recent advancements have extended Tsuji-Wacker oxidation to the of alkaloids, where intramolecular variants promote ring formation in analogs. For instance, in the of the (−)-alstonerine, a modified Wacker oxidation sequence was applied to forge the E ring through selective olefin oxidation, demonstrating compatibility with sensitive motifs. As of 2025, intramolecular Tsuji-Wacker cyclizations continue to enable the construction of fused heterocycles in bioactive alkaloids, such as those featuring or rings, by directing nucleophilic attack on coordinated alkenes. These strategies have been pivotal in synthesizing structurally diverse alkaloids with potential therapeutic properties. The primary advantages of Wacker and Tsuji-Wacker oxidations in synthetic contexts stem from their mild reaction conditions, typically involving aqueous media at ambient temperatures, which tolerate a wide array of s including esters, amides, and heterocycles without requiring harsh oxidants. This functional group compatibility facilitates late-stage oxidations in complex intermediates, minimizing manipulations and preserving molecular integrity during multistep sequences. In , Tsuji-Wacker oxidation is frequently integrated with palladium-catalyzed cross-coupling reactions to support diversity-oriented synthesis, allowing rapid generation of ketone-functionalized libraries from precursors. For example, tandem sequences combining Tsuji-Wacker oxidation with Sonogashira or couplings have been employed to diversify fused γ-lactone scaffolds, yielding compound collections for biological screening. This combinatorial approach leverages shared catalysis to streamline access to structurally varied candidates, enhancing efficiency in lead optimization.

References

  1. [1]
    Hoechst-Wacker Oxidation - an overview | ScienceDirect Topics
    The Wacker process is defined as a catalytic method used to convert alkenes into ketones or aldehydes using a palladium catalyst, ...
  2. [2]
    Wacker-Tsuji Oxidation - Organic Chemistry Portal
    The Wacker Oxidation is an industrial process, which allows the synthesis of ethanal from ethene by palladium-catalyzed oxidation with oxygen.
  3. [3]
    Wacker Oxidation - Chemistry LibreTexts
    Jan 22, 2023 · The Wacker oxidation refers generally to the transformation of a terminal or 1,2-disubstituted alkene to a ketone through the action of catalytic palladium(II) ...
  4. [4]
    Hoechst Reveals Wacker Process Details | C&EN Global Enterprise
    The Wacker process gives high yields of acetaldehyde with a quality comparable to that from the conventional acetylene hydration process.
  5. [5]
    Acetaldehyde from Ethylene—A Retrospective on the Discovery of ...
    Nov 11, 2009 · The Wacker oxidation for the conversion of ethylene into acetaldehyde is one of the most important processes in the chemical industry.Missing: Reinhard | Show results with:Reinhard
  6. [6]
    Cologne, Germany - Wacker Chemie AG
    Our Cologne plant was founded in 1960 as a facility for producing acetylaldehyde from ethylene. Located at the Cologne-Merkenich Chemical Industry Park, ...Missing: first commercial 1959
  7. [7]
    [PDF] A century of expansion and progress
    Feb 23, 2017 · During the 1950s, WACKER began the process of switching from acetylene, to ethylene, feedstock, for its vinyl plastic manufacturing operations.
  8. [8]
    [PDF] People - Markets - Chemistry - Wacker Chemie AG
    Wacker Chemie, founded in 1914, is a pioneer in acetylene-derived chemicals and plastics, and now leads in vinyl polymers, silicones, hyperpure silicon, and ...
  9. [9]
    History - Wacker Chemie AG
    2. The WACKER Process. Basic-research scientists at WACKER based the production of acetylaldehyde and acetic acid on a new raw material. These products could ...
  10. [10]
    Celanese Acetaldehyde Unit on Stream | C&EN Global Enterprise
    Celanese licensed the ethylene-to-acetaldehyde process, known as the Wacker process, from Aldehyd G.m.b.H., which is owned jointly by two German chemical ...
  11. [11]
    Early Industrial Roots of Green Chemistry and the history of the BHC ...
    Oct 25, 2017 · Celanese licensed the Wacker process and opened an acetaldehyde plant in Bay City Texas in 1962, and combined the Wacker acetaldehyde product ...
  12. [12]
    Production of Acetaldehyde via Oxidative Dehydrogenation of ...
    In Europe, route (3), with direct oxidation of ethylene via the Wacker–Hoechst process, dominates because of the lower prices of ethylene compared to ethanol or ...
  13. [13]
    Economic assessment of acetaldehyde production from oxidative ...
    The conventional bioethanol oxidative dehydrogenation to acetaldehyde suffers from low energy efficiency and high capital cost due to cryogenic air ...
  14. [14]
    [PDF] The Wacker Process - Thieme Connect
    The palladium-catalyzed oxidation of ethene to acetaldehyde through addition of water, in the presence of copper(II) chloride and molecular oxygen, ...
  15. [15]
    Stabilization of molecular Pd species in a heterogenized wacker ...
    The active palladium and copper species for the carbonsupported Wacker catalyst were determined to be a molecular Pd(II) species coordinated with chlorine ...
  16. [16]
    Wacker-type oxidation in vapor phase using a palladium–copper ...
    Jan 25, 2009 · Typically, the reaction is carried out in an aqueous solution of PdCl2 and CuCl2 used as co-catalysts. A drawback of this homogeneous catalyst ...
  17. [17]
  18. [18]
    acetaldehyde production, ethylene oxidation - Rest-of-World
    This dataset covers the direct oxidation (Wacker process) of the acetaldehyde production. Namely, ethylene and oxygen, along with a recycle stream, are fed ...
  19. [19]
    [PDF] Wacker Oxidation - stoltz2.caltech.edu
    First organopalladium reaction applied on industrial scale. ♢ First rendered commercial in 1960. ♢ At one point was responsible for the production of over 2 ...
  20. [20]
    [PDF] Intratec.us - Acetaldehyde Production from Ethylene (Two-Step ...
    (2) Acetaldehyde, McKetta, J. J., Encyclopedia of Chemical Processing and Design - Volume 1,. 1976. Keywords: Ethanal, Wacker Process, Hoechst-Wacker Process.
  21. [21]
    Imparting Catalyst-Control upon Classical Palladium-Catalyzed ...
    Nov 23, 2011 · In this Account, we detail our efforts to develop catalyst-controlled variants of both the Wacker oxidation and the Heck reaction to address synthetic ...Missing: precipitation | Show results with:precipitation
  22. [22]
    Cationic palladium(II) complexes as catalysts for the oxidation of ...
    Mar 9, 2015 · Although the original Wacker Process utilised copper salts for the oxidation of ethylene it has been highlighted that copper salts often do ...
  23. [23]
    Strategies to Use O2 as the Oxidant in Organometallic C–H ...
    Jan 23, 2012 · Separately, many recently reported PdII-catalyzed C–H oxidation reactions employ stoichiometric benzoquinone, CuII, or AgI as stoichiometric ...
  24. [24]
    [PDF] Palladium(II)-Catalysed Oxidation of Alkenes - UCL Discovery
    Jul 21, 2015 · Abstract: This review provides a summary of recent developments in the Pd(II)-catalysed oxidation of alkenes,.
  25. [25]
    DFT mechanistic investigation of Wacker-type oxidation of olefins ...
    Oct 2, 2025 · This study investigates the mechanism of Wacker-type oxidation of olefins mediated by the Pd–Quinox catalyst, which incorporates an ...
  26. [26]
    Asymmetric Migratory Tsuji–Wacker Oxidation Enables the ...
    We report herein a generalized enantioselective synthesis of chiral diarylmethanes by means of an asymmetric migratory Tsuji–Wacker oxidation of simple ...
  27. [27]
    The Use of Gases in Flow Synthesis - ACS Publications
    This review will highlight the potential benefits that can be leveraged by using flow chemistry to allow gases to be used in research in a safer and more ...
  28. [28]
    Wacker Oxidation: A Powerful Tool for Natural Product Synthesis
    Apr 10, 2025 · Wacker oxidation is an aerobic oxidation of olefins that has emerged as an active area of research in terms of natural product synthesis.
  29. [29]
    Wacker Oxidation of Terminal Alkenes Over ZrO2‐Supported Pd ...
    Aug 21, 2017 · ZrO2-supported Pd0 NPs with a diameter of 4–12 nm exhibited high catalytic activity for Wacker oxidation and Wacker-type cyclization of terminal ...
  30. [30]
    Palladium/Iron-Catalyzed Wacker-Type Oxidation of Aliphatic ...
    Oct 23, 2023 · The Wacker-type oxidation of aliphatic terminal alkenes proceeds using a Pd/Fe catalyst system under mild reaction conditions using 1 atm O 2 without other ...
  31. [31]
    Acetaldehyde Market Size, Growth, Share, Industry Trends ...
    Aug 11, 2025 · The Acetaldehyde Market is expected to reach 1.19 million tons in 2025 and grow at a CAGR of 2.92% to reach 1.37 million tons by 2030.Missing: 2020s | Show results with:2020s
  32. [32]
    Wacker Up! | Research Communities by Springer Nature
    Feb 28, 2020 · This process, more commonly known as Wacker oxidation, supplies more than 85% of the global acetaldehyde production. To oxidize ethylene ...
  33. [33]
    Acetaldehyde Uses and Applications - Celanese
    Acetaldehyde, which is obtained from the direct oxidation of ethylene, is a pungent, colorless liquid whose boiling point is 21°C, making it possible to be ...Missing: process | Show results with:process
  34. [34]
    [PDF] Acetone Production from Propylene (Single-Step Process) - Intratec.us
    In this process, a catalyst system of palladium chloride and copper chloride is used to oxidize propylene to acetone.
  35. [35]
    Asymmetric Total Synthesis of Taxol® By Teruaki MUKAIYAMA ...
    wacker oxidation conditions. By the above sequence of manipulations, the ABC ring system precursor 23 was efficiently synthesized from optically active 8 ...
  36. [36]
    Next-Generation Total Synthesis of Vancomycin - ResearchGate
    Sulfamate-tethered aza-Wacker cyclization followed by a Tsuji-Wacker ketone synthesis furnishes a key tricyclic intermediate which we hypothesize can be ...
  37. [37]
    Total synthesis of selected bioactive alkaloids, their structure ...
    Liao & coworkers have developed a modified Wacker oxidation sequence to get E ring for the total synthesis of (−)-alstonerine.[31] The intermediate 16-epi-Na- ...
  38. [38]
    Peroxide-Mediated Wacker Oxidations for Organic Synthesis - PMC
    Oct 12, 2016 · b) The reaction is proposed to occur via alcohol oxidation to provide a Pd–H, into which molecular oxygen can insert. ... 10 mol % PdCl2, 1 equiv.
  39. [39]
    PdII Catalysis: Recent Advances in the Intramolecular Wacker-Type ...
    Tsuji, J. Synthetic applications of the palladium-catalyzed oxidation of olefins to ketones. Synthesis 1984, 1984, 369–384. [Google Scholar] [CrossRef] ...
  40. [40]
    Impact of Cross-Coupling Reactions in Drug Discovery and ... - MDPI
    Cross-coupling reactions have made an impressive impact in medicinal chemistry, and drug discovery and development for more than two decades.