Fact-checked by Grok 2 weeks ago

Acetoxy group

The acetoxy group, also known as the acetyloxy group, is a in characterized by the −O−C(=O)−CH₃, where an oxygen atom is bonded to the carbonyl carbon of an acetyl (CH₃CO−) unit. This group functions as an linkage and is commonly abbreviated as AcO or OAc in chemical notation. In IUPAC substitutive nomenclature, the preferred prefix is "acetyloxy" for the CH₃−CO−O−, although the contracted form "acetoxy" is permitted and frequently used in general literature and practice. The acetoxy group plays a central role in and natural products, often serving as a for hydroxyl (−OH) functionalities on alcohols or to prevent unwanted reactivity during multi-step reactions. It is typically introduced through using reagents like (Ac₂O) in the presence of a base such as , and deprotected via under acidic or basic conditions, or enzymatically with lipases for selective removal. In pharmaceuticals, the acetoxy group is prominent in acetylsalicylic acid (aspirin), where it is attached ortho to a on a ring, enhancing and enabling its and effects by acting as a that acetylates enzymes. Beyond , acetoxy-substituted compounds are key in ; for instance, (CH₂=CH−OCOCH₃) contains this group and serves as a for producing , a widely used in , paints, coatings, and textiles due to its adhesive properties and film-forming ability. Acetoxylation reactions, which install the group via C−H activation or often catalyzed by or salts, are valuable for synthesizing complex molecules in agrochemicals, antimicrobials, and herbicides.

Definition and Structure

Chemical Composition

The acetoxy group is a in with the formula −OC(O)CH₃, consisting of an oxygen atom bonded to a carbonyl carbon that is double-bonded to another oxygen atom and single-bonded to a (CH₃). According to IUPAC nomenclature, it is systematically named "acetyloxy" when used as a , with "acetoxy" as the retained and preferred form in general nomenclature. In molecular structures, the acetoxy group is represented as R−OC(=O)−CH₃, where R denotes the parent chain or to which the group is attached, forming an linkage. This group is derived from acetic acid (CH₃COOH) through the removal of the hydroxyl , resulting in the characteristic connectivity. In the context of esters, it is traditionally referred to as the group, a naming practice that emerged in early 19th-century literature following the introduction of the term "" by Leopold Gmelin in 1848.

Molecular Geometry

The carbonyl carbon in the acetoxy group (-OC(O)CH₃) is sp² hybridized, resulting in a around the C=O moiety with bond angles of approximately 120°. This hybridization facilitates the overlap of the carbon's p orbital with the oxygen's p orbital, forming the π bond characteristic of the carbonyl. Resonance stabilization plays a key role in the electronic structure of the acetoxy group, where the on the ester oxygen donates electron density into the carbonyl π* antibonding orbital, delocalizing the π electrons across the O-C=O system. This delocalization imparts partial double-bond character to the ester C-O bond, shortening it to about 1.36 Å—shorter than a typical C-O (1.43 Å)—while slightly lengthening the C=O bond compared to isolated carbonyls. Computed bond lengths from (MM4) for the acetoxy group in confirm these features: the C=O bond is approximately 1.21 , the ester C-O bond is 1.36 , and the C-CH₃ bond is 1.50 . These values reflect the balance between σ and π bonding influenced by . Due to this conjugation, the acetoxy group enforces planarity on adjacent atoms in the parent molecule, aligning the ester framework to maximize orbital overlap and stabilize the overall structure.

Physical and Chemical Properties

Solubility and Stability

The acetoxy group (-OCOCH₃) imparts to compounds through its carbonyl functionality, thereby enhancing their in polar solvents such as , acetone, and . For instance, , a representative , is miscible with , acetone, and . This arises from the ability of the to participate in dipole-dipole interactions and hydrogen bonding with these solvents. In contrast, esters exhibit limited in due to their hydrophobic alkyl components; , for example, dissolves to approximately 8 g/100 mL at 20 °C, while longer-chain alkyl acetates are generally insoluble unless the parent chain includes hydrophilic moieties that promote aqueous interactions. Acetate esters demonstrate good thermal stability under ambient conditions, remaining intact up to temperatures of 200–250 °C in many applications, particularly in polymeric contexts where the acetoxy group contributes to structural integrity. However, at higher temperatures, can occur via pathways such as β-elimination in β-acetoxy-substituted compounds, leading to the release of acetic acid and formation of alkenes. For simple esters like , is minimal at 400 °C, with significant breakdown requiring temperatures exceeding 500 °C, often producing acetic acid, , and other fragments. Chemically, compounds bearing the acetoxy group are generally resistant to mild conditions, as the linkage withstands non-nucleophilic bases without significant reaction. This stems from the low nucleophilicity required to attack the carbonyl carbon under or weakly environments. However, they are susceptible to acid-catalyzed , where of the carbonyl oxygen facilitates nucleophilic attack by water, yielding the corresponding and acetic ; the pKₐ of this protonated carbonyl conjugate is approximately -7, indicating strong acidity and ease of protonation in acidic media. Brief exposure to hydrolysis conditions highlights a limit, with full details on the covered elsewhere. Representative examples illustrate these properties: , with a boiling point of 72 °C, is volatile and remains stable in air when properly inhibited against , making it suitable for industrial handling at moderate temperatures. In comparison, alkyl acetates such as (boiling point 77 °C) exhibit greater thermal endurance without polymerization risks and similar profiles, though they decompose more readily under prolonged heating above 400 °C.

Spectroscopic Characteristics

The acetoxy group, characteristic of esters, exhibits distinct () bands that facilitate its identification. The carbonyl (C=O) stretching vibration appears as a strong in the range of 1730–1750 cm⁻¹ for aliphatic acetates, reflecting the conjugated nature of the functionality. Additionally, the C–O stretching vibration manifests as a strong band between 1200 and 1300 cm⁻¹, often more pronounced around 1240 cm⁻¹ in esters due to the asymmetric C–C–O mode. The absence of a broad O–H stretching band (typically 3200–3600 cm⁻¹) in the spectrum of an acetoxylated compound confirms successful formation from the corresponding . In (NMR) , the acetoxy group produces characteristic signals in both ¹H and ¹³C spectra. The methyl protons (CH₃) of the acetyl moiety appear as a sharp at 1.9–2.1 ppm in ¹H NMR, deshielded by the adjacent carbonyl, as observed in simple acetate esters like . In ¹³C NMR, the carbonyl carbon resonates at 170–175 ppm, indicative of the environment, while the methyl carbon shifts to 20–22 ppm, providing clear markers for the acetoxy substituent./Spectroscopy/Magnetic_Resonance_Spectroscopies/Nuclear_Magnetic_Resonance/NMR%3A_Structural_Assignment/Interpreting_C-13_NMR_Spectra) Electron ionization mass spectrometry (EI-MS) of compounds bearing the acetoxy group often reveals diagnostic fragment ions arising from characteristic cleavage patterns. A prominent peak at m/z 43 corresponds to the acetyl cation (CH₃CO⁺), formed via alpha-cleavage adjacent to the carbonyl. Another common ion at m/z 60 arises from the molecular ion of acetic acid (CH₃COOH⁺•), resulting from a McLafferty rearrangement involving the ester oxygen and a gamma-hydrogen if available in the alkyl chain. Ultraviolet-visible (UV-Vis) spectroscopy of the acetoxy group shows weak absorption around 200–220 nm, attributed to the π–π* transition in the carbonyl chromophore, with the intensity and exact position influenced by the molecular environment. This feature is particularly useful for detecting acetoxy groups in conjugated systems, though simple aliphatic acetates exhibit only end absorption below 220 nm.

Synthesis Methods

Esterification of Alcohols

The acetoxy group is commonly introduced to alcohols through esterification reactions, with the esterification serving as a foundational method for synthesizing acetate esters in laboratory settings. This process involves the acid-catalyzed condensation of an (ROH) with acetic acid (CH₃COOH), resulting in the formation of the acetate ester (R-OC(O)CH₃) and as a byproduct. The reaction is reversible and reaches equilibrium, governed by , where excess reactants can shift the equilibrium toward product formation. The general equation for Fischer esterification of alcohols to acetates is: \text{ROH} + \text{CH}_3\text{COOH} \rightleftharpoons \text{R-OC(O)CH}_3 + \text{H}_2\text{O} Typically, the reaction employs a strong acid catalyst such as concentrated (H₂SO₄) to protonate the carbonyl oxygen of acetic acid, facilitating nucleophilic attack by the . Conditions often involve refluxing the mixture in excess acetic acid for 2–24 hours, depending on the 's reactivity, followed by extraction and distillation to isolate the ester. For primary , yields generally range from 70% to 95%, though equilibrium limitations may require removal of water (e.g., via Dean-Stark apparatus) or use of excess acetic acid to improve efficiency. This method, first systematically described by and Arthur Speier in 1895, has been widely applied since the late for preparing simple acetate esters like . A widely used method for the of alcohols involves reaction with ((CH₃CO)₂O) in the presence of a . The acts as a , attacking one carbonyl carbon of the anhydride to form the acetate ester and acetic acid as a . A such as or triethylamine neutralizes the acetic acid produced. The general equation is: \text{ROH} + (\text{CH}_3\text{CO})_2\text{O} \rightarrow \text{R-OC(O)CH}_3 + \text{CH}_3\text{COOH} This approach typically proceeds under mild conditions, often at in an inert like , and provides high yields (often >90%) with short reaction times (minutes to hours). It is particularly favored for introducing the acetoxy group as a due to its efficiency, selectivity, and compatibility with acid-sensitive substrates. An alternative and more reactive approach utilizes (CH₃COCl) for the direct of , offering faster reaction times and higher yields compared to Fischer esterification. The reaction proceeds via , where the attacks the carbonyl carbon of , displacing chloride to form the and HCl. A base such as or triethylamine is commonly added to neutralize the HCl and prevent side . The equation is: \text{ROH} + \text{CH}_3\text{COCl} \rightarrow \text{R-OC(O)CH}_3 + \text{HCl} This method is particularly suitable for sensitive alcohols, achieving near-quantitative yields under mild conditions (often at room temperature in an inert solvent like dichloromethane) and completing in minutes to hours, making it preferable for scale-up or when avoiding harsh acids is necessary.

Acetylation of Carboxylates

The acetylation of carboxylates provides an important route for synthesizing mixed carboxylic anhydrides containing the acetoxy group, of the general form R-C(O)-O-C(O)-CH3, where R represents an alkyl or aryl substituent from the original carboxylic acid. These compounds serve as activated derivatives of carboxylic acids, facilitating subsequent nucleophilic acyl substitution reactions such as amide bond formation in peptide synthesis or esterification. The reaction proceeds via nucleophilic attack by the carboxylate oxygen on the carbonyl carbon of acetic anhydride, displacing acetate. A primary involves the direct of a with , as illustrated by the equation RCOOH + (CH_3CO)_2O → RCO-OC(O)CH_3 + CH_3COOH. This process favors the mixed anhydride under appropriate conditions, with the first step being second-order and an of approximately 80 kJ/mol. For instance, reacts with at 30–70 °C in a 1:1 ratio to yield the acetic-oleic mixed anhydride at , reaching completion in about 90 minutes without a catalyst. The for this step ranges from 2.21 to 2.57, indicating moderate favorability toward the product. An alternative approach utilizes salts, such as sodium carboxylates, reacting with according to RCOO^- Na^+ + (CH_3CO)_2O → RCO-OC(O)CH_3 + CH_3COONa. This variant is particularly prevalent in industrial settings for preparing mixed anhydrides of fatty acids, leveraging the higher nucleophilicity of the carboxylate anion to drive the efficiently. The process typically occurs at , offering high yields and circumventing the water sensitivity inherent in methods involving free alcohols. These mixed anhydrides are distinct from simple alkyl acetates, as they derive from precursors and exhibit anhydride-specific reactivity.

Reactivity and Reactions

Hydrolysis and Deprotection

The hydrolysis of the acetoxy group, -OC(O)CH₃, cleaves the ester linkage to regenerate the parent alcohol and acetic acid (or acetate), serving as a key deprotection step in organic synthesis. This reaction proceeds via either acid- or base-catalyzed mechanisms, with the choice depending on the substrate's sensitivity and the presence of orthogonal protecting groups. In acid-catalyzed hydrolysis, the acetoxy ester R-OC(O)CH₃ reacts with water in the presence of hydronium ion (H₃O⁺) to yield ROH + CH₃COOH. Typical catalysts include HCl or H₂SO₄, and the reaction rate increases with decreasing pH and rising temperature; for instance, under strongly acidic conditions at elevated temperatures, hydrolysis can achieve completion within hours. Base-catalyzed hydrolysis, known as saponification, involves nucleophilic attack by hydroxide ion on the carbonyl carbon: R-OC(O)CH₃ + OH⁻ → ROH + CH₃COO⁻. This process is generally faster for simple acetate esters compared to more sterically hindered or electronically deactivated esters, allowing selectivity in complex molecules where acetates are preferentially cleaved over other ester types. For deprotection purposes, mild conditions such as treatment with aqueous acetic acid enable the regeneration of alcohols from acetoxy groups without disrupting sensitive functionalities. This approach exhibits orthogonality to silyl protecting groups, which remain intact under these weakly acidic conditions, facilitating selective manipulation in multi-step syntheses. Enzymatic hydrolysis using lipases, such as Candida antarctica lipase B, provides a selective for deprotecting acetoxy groups under mild conditions, often in solvents or aqueous , preserving other sensitive functionalities. This biocatalytic approach is particularly valuable in pharmaceutical and synthesis for regioselective removal. The kinetics of base-catalyzed acetate ester hydrolysis are second-order overall ( in ester and in ), with a representative second-order rate constant of approximately 0.11 M⁻¹ s⁻¹ at 25°C for .

Transesterification

Transesterification involving the acetoxy group refers to the exchange of the alkoxy substituent in an acetate ester (R-OC(O)CH₃) with the hydroxyl group of another alcohol (R'OH), resulting in a new ester (R'-OC(O)CH₃) and the release of the original alcohol (ROH). This reaction proceeds through a nucleophilic acyl substitution mechanism, which can be catalyzed by either acids or bases. In the acid-catalyzed pathway, protonation of the carbonyl oxygen enhances the electrophilicity of the carbonyl carbon, facilitating nucleophilic attack by the alcohol, followed by proton transfers and elimination of the leaving group alcohol. The base-catalyzed mechanism involves deprotonation of the attacking alcohol to form an alkoxide nucleophile, which adds to the carbonyl, leading to a tetrahedral intermediate and subsequent expulsion of the acetate-leaving alcohol. The reaction is reversible, with the equilibrium governed by the relative stabilities of the esters and alcohols involved; it can be driven forward by employing an excess of the desired alcohol or by continuously removing the byproduct alcohol, such as through distillation. Common catalysts for acetoxy group transesterification include strong acids like (H₂SO₄) for non-selective processes and enzymes such as lipases for regioselective applications, particularly in the of complex s. Acid-catalyzed reactions typically occur under conditions with the , achieving yields ranging from 50% to 90% depending on the substrates and optimization, as demonstrated in the conversion of to using ionic liquids or resin catalysts. Lipases, often immobilized for reusability, enable mild, solvent-free or organic media conditions at ambient temperatures, favoring the use of activated acetate donors like for efficient acyl transfer in asymmetric . These enzymatic methods are particularly valuable in natural product chemistry, where allows selective of primary over secondary hydroxyl groups in polyols. Acetoxy groups exhibit enhanced lability in compared to esters with longer acyl chains, primarily due to reduced steric hindrance around the small in the acetyl moiety, which facilitates nucleophilic approach and departure. This reactivity makes acetate esters preferred acyl donors in both chemical and enzymatic processes. Industrially, with acetates finds application in , where waste cooking oils react with under supercritical conditions or with acid catalysts to yield ethyl esters alongside byproducts, offering a glycerol-free alternative to traditional methanolysis with yields up to 95%.

Applications in Chemistry

Protecting Groups for Alcohols

The acetoxy group, derived from of , serves as a temporary in to mask hydroxyl functionalities and prevent unwanted side reactions, such as oxidation or interference during processes./15%3A_Alcohols_and_Ethers/15.10%3A_Protection_of_Hydroxyl_Groups) This protection is particularly valuable in multi-step syntheses of complex molecules, where the alcohol's reactivity could otherwise complicate subsequent transformations. The acetoxy group is typically introduced by treating the alcohol with (Ac₂O) in the presence of , a mild that facilitates the esterification while minimizing side reactions. Key advantages of the acetoxy include its low cost, straightforward installation using readily available , and ease of removal under mild conditions, making it suitable for routine applications. It exhibits good compatibility with a variety of , including those used in oxidation and reduction reactions, though it is less stable under basic conditions compared to silyl ethers like tert-butyldimethylsilyl (), which resist base-mediated cleavage. However, limitations arise from its tendency to undergo acyl migration under acidic conditions, particularly in vicinal diols where the ester can shift between adjacent oxygen atoms, potentially disrupting . Additionally, it is not ideal for long-term protection in prolonged syntheses due to susceptibility to or migration over time. In chemistry, the acetoxy group enables selective of primary alcohols over secondary ones, facilitating targeted modifications in polyhydroxylated systems. For instance, of primary-secondary diols with in the presence of neutral alumina yields the corresponding primary monoacetates in good yields without forming diacetates, allowing precise control in assembly. Similarly, in the of and related glycopeptide antibiotics, acetoxy groups have been employed to protect secondary alcohols during steps, as demonstrated in approaches by Nicolaou and coworkers, where acetate-protected disaccharides improved coupling efficiency and stereoselectivity in late-stage assembly. Deprotection of these groups can be achieved through methods such as , as detailed in the reactivity section.

Role in Polymer Chemistry

The acetoxy group plays a pivotal role in through its incorporation into , a key that undergoes to form (PVAc). This process typically yields polymers with molecular weights ranging from approximately 100,000 to 500,000 g/mol, enabling the production of versatile materials such as adhesives for wood and paper, as well as coatings for architectural paints and primers. A significant application of PVAc involves its partial , or , to produce (PVOH), where acetoxy groups are replaced by hydroxyl groups. This transformation results in PVOH with a temperature (Tg) of approximately 85°C, making it suitable for textiles, where it provides enhanced oil, grease, and wear resistance in manufacturing processes. In addition, the acetoxy group is central to the synthesis of via the of , typically achieving a degree of substitution () of 2 to 3 acetoxy groups per glucose unit. This derivatization imparts in acetone, facilitating the formation of films for and , as well as fibers for textiles and filters. On an industrial scale, the global production of , the primary precursor for PVAc and related polymers, exceeded 5 million tons per year in the , underscoring its importance in sectors like paints, adhesives, and paper processing.

References

  1. [1]
    Blue Book P-60-65 - IUPAC nomenclature
    Compounds containing the –N=C=O group attached to a parent hydride structure, are named by using substitutive nomenclature and the prefix 'isocyanato'. This ...
  2. [2]
    [PDF] Protective Groups - Harvard
    Several methods for forming and cleaving acetate esters have been developed. Lipases can often be used for the enantioselective hydrolysis of acetate esters ...
  3. [3]
    Aspirin | C9H8O4 | CID 2244 - PubChem - NIH
    Aspirin | C9H8O4 | CID 2244 - structure, chemical names, physical and ... hydroxy group has been replaced by an acetoxy group. A non-steroidal anti ...
  4. [4]
    Vinyl Acetate | C4H6O2 | CID 7904 - PubChem
    Vinyl acetate's production and use as a monomer for making poly(vinyl acetate) and vinyl acetate copolymers, in the production of paints, sealants, coatings ...
  5. [5]
    Cu(OAc)2.H2O used as acetoxy source for the selective C(sp2)‐H ...
    Mar 5, 2025 · Consequently, acetoxy groups play a significant role in pharmaceuticals, contributing to pro-drugs, antimicrobials, and herbicides. They are ...
  6. [6]
    [PDF] Brief Guide to the Nomenclature of Organic Chemistry - IUPAC
    For naming purposes, a chemical compound is treated as a combination of a parent compound. (Section 5) and characteristic (functional) groups, one of which is.
  7. [7]
    What Is an Ester in Chemistry? - ThoughtCo
    Jun 9, 2025 · The term "ester" was coined by the German chemist Leopold Gmelin in 1848. It is likely the term was a contraction of the German word "essigä ...
  8. [8]
    Ethyl Acetate
    Summary of each segment:
  9. [9]
    Thermal Properties and Crystallization Behavior of Curdlan Acetate ...
    Jan 17, 2024 · Curdlan triacetate (CDTAc) has a high glass-transition temperature (171 °C) and melting point (287 °C), but its 50% thermal decomposition ...Missing: stability | Show results with:stability
  10. [10]
    Experimental and theoretical study of the thermal decomposition of ...
    At 400 °C no peaks of ethylene and acetic acid can be observed in the chromatograms (Figure SM3), showing that the conversion of ethyl acetate is nearly zero.
  11. [11]
    Thermal Decomposition of Potential Ester Biofuels. Part I: Methyl ...
    An important goal of this paper is to identify all pathways for the thermal cracking of methyl acetate and methyl butanoate at pressures of roughly 20 Torr and ...Missing: stability | Show results with:stability
  12. [12]
  13. [13]
    [PDF] values to know 1. Protonated carbonyl pKa
    Approximate pKa chart of the functional groups: values to know. 1. Protonated carbonyl pKa = -7 Other important pKa's. 2. Protonated alcohol or ether pKa ...
  14. [14]
    Vinyl Acetate
    Summary of each segment:
  15. [15]
    IR Spectroscopy Tutorial: Esters
    The carbonyl stretch C=O of aliphatic esters appears from 1750-1735 cm-1; that of α, β-unsaturated esters appears from 1730-1715 cm-1. See also: ... The C–O ...
  16. [16]
    Pendant Ester Polymers and Polycarbonates | Spectroscopy Online
    Dec 1, 2022 · Recall that acetate esters have a special high wavenumber C-C-O stretching peak at ~1240 cm -1 (going forward, assume all peak positions are in ...
  17. [17]
    Infrared Spectroscopy Absorption Table - Chemistry LibreTexts
    Sep 11, 2025 · The following table lists infrared spectroscopy absorptions by frequency regions. 4000-3000 cm -1 3000-2500 cm -1 2400-2000 cm -1 2000-1650 cm -1
  18. [18]
    Esters
    In esters, these protons are shifted to 2-2.2 ppm. The protons on the “other” side of the ester, meaning those attached to the carbon adjacent to a carbonyl ...Missing: 13C | Show results with:13C
  19. [19]
    (PDF) UV absorption cross sections for acetates - ResearchGate
    Aug 9, 2025 · The UV absorptions were measured at the wavelength range of 200 to 260 nm for methyl and 200–329 nm for acetate at room temperature.
  20. [20]
    Kinetic Analysis as an Optimization Tool for Catalytic Esterification ...
    Apr 30, 2020 · In synthesis, the Brønsted acid-catalyzed Fischer esterification reaction has served the community well since 1895. (1) However, this ...
  21. [21]
    [PDF] 5.310 (F19) Fischer Esterification Lab Manual - MIT OpenCourseWare
    Esters are made, by condensing an alcohol with a carboxylic acid. The reaction is generally catalyzed with concentrated sulfuric acid (H2SO4).
  22. [22]
    [PDF] Fischer Esterification
    A typical procedure to synthesize esters is the Fischer esterification, wherein a carboxylic acid is treated with an alcohol in the presence of a mineral ...
  23. [23]
    [PDF] 100 Chapter 21. Carboxylic Acid Derivatives and Nucleophilic Acyl ...
    Alcoholysis: Acid chlorides react with alcohols to give esters. reactivity: 1° alcohols react faster than 2° alcohols, which react faster than 3° alcohols. R.
  24. [24]
  25. [25]
    Acetic Anhydride and Mixed Fatty Acid Anhydrides - ResearchGate
    Acetic anhydride production through acetic acid represents, along with the ketene process, the main commercial means of acetic anhydride production. 136,161 ...
  26. [26]
    Acid-catalyzed hydrolysis of phenyl acetate - ACS Publications
    Acid-catalyzed hydrolysis of phenyl acetate ... Rodríguez-Dafonte. Change in the Acid Hydrolysis Mechanism of Esters Enforced by Strongly Acid Microemulsions.
  27. [27]
    Ester Hydrolysis: Acid and Base-Catalyzed Mechanism
    Ester hydrolysis mechanism can be acid or base catalyzed. Base-catalyzed ester hydrolysis has the advantage of being irreversible.
  28. [28]
    Acetic Acid Esters - Organic Chemistry Portal
    Various alcohols, thiols, phenols, and amines can be acetylated using acetic anhydride in the presence of catalytic quantity of silver triflate.Missing: acetoxy | Show results with:acetoxy
  29. [29]
    Silyl-protective groups influencing the reactivity and selectivity ... - NIH
    Jan 16, 2017 · In carbohydrate chemistry silyl protective groups have frequently been used primarily as an orthogonal protective group to the more commonly ...
  30. [30]
    Transesterification - Master Organic Chemistry
    Nov 10, 2022 · Transesterification is the conversion of one ester into another through exchange of -OR groups, and can occur under basic or acidic conditions.Missing: acetoxy | Show results with:acetoxy
  31. [31]
    Transesterification - Chemistry LibreTexts
    Jan 22, 2023 · Transesterification is the conversion of a carboxylic acid ester into a different carboxylic acid ester, involving an exchange of alkoxy groups.Missing: acetoxy | Show results with:acetoxy
  32. [32]
    The mechanism and thermodynamics of transesterification of ...
    In solution, base-catalyzed hydrolysis and transesterification of esters are initiated by hydroxide- or alkoxide-ion attack at the carbonyl carbon.
  33. [33]
    Kinetics of Dowex 50W catalyzed transesterification of methyl ...
    Feb 6, 2024 · The esters produced in this investigation (n-propyl, n-butyl, and isobutyl acetates) are very important and can be used in many applications ...
  34. [34]
    Lipase catalyzed synthesis of cinnamyl acetate via ...
    This work focuses on the synthesis of cinnamyl acetate via lipase catalyzed transesterification of cinnamyl alcohol with vinyl acetate in non-aqueous medium.
  35. [35]
    A review on enzymatic synthesis of aromatic esters used as flavor ...
    The present article gives an overview of the aromatic esters synthesis, considering the main effects in the reaction media conditions and enzymes used.Review · Abstract · Introduction
  36. [36]
    Lipase-catalyzed irreversible transesterification for preparative ...
    ... Lipase Catalyzed Transesterification with Isopropenyl Acetate in Organic Solvent. Journal of Oleo Science 2005, 54 (2) , 105-114. https://doi.org/10.5650 ...
  37. [37]
    Relative Rates of Transesterification vis-à-vis Newman's Rule of Six
    Sep 6, 2024 · We have systematically examined the relative rates of base-catalyzed transesterification for ten model acetates containing, in total, a ...
  38. [38]
    From waste cooking oil and ethyl acetate by using supercritical ...
    Nov 19, 2024 · This study examines the transesterification method for producing biodiesel from two sources: leftover cooking oil and ethyl acetate.
  39. [39]
    Glycerol free biodiesel synthesis by application of methyl formate in ...
    May 23, 2023 · The process of interesterification, like the process of transesterification, takes place in three stages. In the first, a molecule of ...<|separator|>
  40. [40]
    O-Acetylation using acetic anhydride in pyridine - NCBI - NIH
    Oct 6, 2021 · Introduction. The acetyl (Ac) group is one of the most commonly used protecting groups in carbohydrate chemistry (1).
  41. [41]
    Protecting group migrations in carbohydrate chemistry - ScienceDirect
    Migration of commonly used groups like silyl, acetal and acyl groups under various reaction conditions are discussed.
  42. [42]
    The selective acetylation of primary alcohols in the presence of ...
    Abstract. Treatment of primary-secondary sugar diols with ethyl acetate in the presence of Woelm neutral alumina produced selectively the corresponding primary ...
  43. [43]
    Total Syntheses of Vancomycin Related Glycopeptide Antibiotics ...
    A review of efforts that have provided total syntheses of vancomycin and related glycopeptide antibiotics, their agylcons, and key analogues is provided.
  44. [44]
    Vinyl acetate - American Chemical Society
    Oct 19, 2020 · Vinyl acetate is an important industrial monomer that is used to make homopolymers and copolymers with a wide variety of applications.Missing: group | Show results with:group
  45. [45]
    Vinyl Acetate Monomer (VAM): A Highly Versatile Polymerization ...
    The largest end-use for VAM is in the production of polyvinyl acetate resins as a base for adhesives and coatings, as well as a feedstock for derivative resins ...
  46. [46]
    An Introduction to Vinyl Acetate-Based Polymers
    Vinyl acetate monomers (VAM) are essential building blocks for a large number of water-based polymers. Vinyl acetate is prepared from ethylene by reacting it ...Missing: acetoxy | Show results with:acetoxy
  47. [47]
    [PDF] polyvinyl-alcohol-a-comprehensive-overview.pdf
    Hydrolysis: The polyvinyl acetate is then hydrolyzed to convert the acetate groups to hydroxyl groups. This can be accomplished using aqueous alkali, such as ...
  48. [48]
    Why Polyvinyl Alcohol (PVA) Is Essential in Modern Polymer ...
    Instead, PVA is produced through the hydrolysis of polyvinyl acetate (PVAc) ... Textile and paper manufacturing: PVA enhances oil, grease, and wear ...
  49. [49]
    High-Performance Acetylated Ioncell-F Fibers with Low Degree of ...
    May 30, 2018 · It is demonstrated that cellulose acetate can be synthesized with different degree of substitution (DS) values, and that some commonly used ...
  50. [50]
    Synthesis and Characterization of Cellulose Triacetate Obtained ...
    Feb 21, 2022 · ... acetone-soluble cellulose acetate with a degree of substitution from 2.5–2.8 [8]. Similarly, sugarcane bagasse was used to isolate pure ...
  51. [51]
    Worldwide Vinyl Acetate Monomer Industry to 2030 - Business Wire
    Feb 2, 2022 · The global vinyl acetate monomer demand stood at 5.2 Million Tonnes in 2020 and is forecast to reach 9.097 Million Tonnes by 2030, growing ...Missing: 2020s | Show results with:2020s