Fact-checked by Grok 2 weeks ago

Transesterification

Transesterification is a in which the alkoxy group of an is exchanged for the alkoxy group of an , yielding a new and a new . This process, also known as alcoholysis, is reversible and typically requires a catalyst such as an , , or to proceed efficiently, with the often shifted toward the products by using an excess of the alcohol reactant in accordance with . The mechanism of transesterification generally follows an addition-elimination pathway. In base-catalyzed transesterification, an alkoxide ion acts as a to attack the carbonyl carbon of the , forming a tetrahedral intermediate, followed by the elimination of the original alkoxy group to form the new . Acid-catalyzed versions involve of the carbonyl oxygen to enhance electrophilicity, of the , proton transfers, and eventual elimination of the leaving group, often described by the sequence protonation-addition-deprotonation-protonation-elimination-deprotonation (PADPED). Enzymatic transesterification, using lipases, offers milder conditions and higher selectivity, particularly for substrates. Transesterification plays a pivotal role in industrial applications, most notably in the production of , where vegetable oils or animal fats (triglycerides) react with in the presence of a base catalyst like to yield methyl esters () and as a . This process converts renewable feedstocks into a viable ; global production of reached approximately 44 million metric tons in 2023. Beyond biofuels, transesterification is essential in for interconverting esters to tailor solubility, reactivity, or physical properties, such as forming cyclic esters (lactones) in polymer precursors or pharmaceutical intermediates. It also finds use in the synthesis of carbonates from and , serving as green solvents or chemical building blocks in sustainable chemistry.

Fundamentals

Definition

Esters are organic compounds derived from carboxylic acids, characterized by the general structure R-COO-R', where R and R' are alkyl or aryl groups, formed by replacing the hydroxyl group (-OH) of the carboxylic acid with an alkoxy group (-OR'). Transesterification is a chemical reaction in which an ester (RCOOR') reacts with an alcohol (R''OH) to produce a new ester (RCOOR'') and a different alcohol (R'OH), typically catalyzed by acids, bases, or enzymes. This reaction plays a crucial role in for interconverting esters, in for modifying structures, and in production for converting triglycerides into fatty acid alkyl esters like . It allows modification of ester functionalities while preserving the carbonyl bond, enabling efficient transformations without . Transesterification is a reversible with an near unity, governed by ; the reaction can be driven toward product formation by using excess alcohol or removing byproducts, such as in synthesis.

General Reaction

The general transesterification reaction involves the exchange of alkoxy groups between and , yielding a new ester and a new . This is depicted by the balanced equation: \ce{RCO2R' + R''OH ⇌ RCO2R'' + R'OH} Here, \ce{RCO2R'}\ ) represents the original ester, \(\ce{R''OH} is the reacting , \ce{RCO2R''} is the product ester, and \ce{R'OH} is the displaced . The reaction is reversible and reaches , characterized by the K, expressed as: K = \frac{[\ce{RCO2R''}][\ce{R'OH}]}{[\ce{RCO2R'}][\ce{R''OH}]} This constant reflects the ratio of product to reactant concentrations at and is influenced by as well as the relative thermodynamic stabilities of the esters involved, with more stable esters favoring higher K values under similar conditions. To favor product formation and shift the equilibrium rightward per , strategies include employing an excess of the reacting or continuously removing a , such as via . In industrial , for instance, excess is used as the , and unreacted is often recovered by to enhance efficiency and drive conversion. Transesterification relates to esterification, the direct formation of an from a and , and , the cleavage of an by to yield a and ; it positions as a specialized exchange where an , rather than , serves as the , maintaining similar reversibility but altering product outcomes.

Catalysts and Conditions

Types of Catalysts

Transesterification reactions commonly employ , , enzymatic, and emerging heterogeneous catalysts, each activating the ester or differently to facilitate the exchange of alkoxy groups. catalysts, such as (H₂SO₄) and (HCl), protonate the carbonyl oxygen of the , enhancing its electrophilicity and enabling nucleophilic attack by the . These homogeneous catalysts tolerate high free fatty acid (FFA) content in feedstocks, avoiding side reactions like , but require careful handling due to equipment corrosion and difficulty in recovery. Base catalysts, including (NaOH) and (KOH), operate by deprotonating the to generate a more nucleophilic ion that attacks the carbonyl, offering faster particularly for transesterification in low-FFA oils. However, their to FFAs leads to formation, limiting use to refined feedstocks with less than 0.5% FFA, and they generate during neutralization. Enzymatic catalysts, primarily lipases such as antarctica lipase B, catalyze the through a serine-based that facilitates acyl-enzyme intermediate formation, providing high selectivity and enantioselectivity for chiral syntheses under mild conditions. These biocatalysts excel in tolerating FFAs and water, enabling processing of waste oils, but their high cost—often exceeding $100/kg for immobilized forms—and slower rates compared to chemical catalysts restrict widespread adoption. Emerging catalysts address limitations of homogeneous systems by promoting for easier separation and reuse in industrial applications. Solid bases like (CaO) supported on materials such as alumina or silica provide active sites for formation while reducing and corrosiveness, enhancing in processes like . Ionic liquids, functioning as both solvents and catalysts with tunable acid-base properties (e.g., [BMIM][HSO₄]), offer advantages through recyclability and high yields with diverse substrates, though their expense and scalability challenges persist. Selection of catalysts depends on reaction medium, with homogeneous types favored for simplicity and speed in batch processes, while heterogeneous variants prioritize reusability and reduced environmental impact in continuous operations. tolerance guides choices, as and enzymatic catalysts handle impure feedstocks better than bases, and cost considerations often tip toward inexpensive homogeneous s or bases for large-scale use despite separation drawbacks.

Reaction Conditions

Transesterification reactions are typically conducted under mild conditions to favor the exchange of alkoxy groups while minimizing side reactions. For base-catalyzed processes, temperatures range from 50°C to 150°C, depending on the substrates and catalyst, with common ranges of 50–65°C for methanolysis using catalysts like NaOH or KOH. Enzymatic transesterification, employing lipases, operates at lower temperatures of 20–60°C to preserve enzyme activity, often around 30–50°C for from oils. Most reactions proceed at , though supercritical conditions using alcohols like at 200–400°C and elevated pressures (e.g., 8–20 ) enable catalyst-free processes for enhanced rates. Solvents play a critical role in facilitating phase mixing and reaction efficiency, with excess frequently serving dual roles as reactant and ; for instance, is commonly used in at molar excesses to solubilize triglycerides. In cases involving immiscible or viscous feedstocks, inert s such as may be added to improve solubility and homogeneity without participating in the reaction. must be rigorously excluded, as even trace amounts promote of esters to carboxylic acids and alcohols, reducing yields and complicating purification. To shift the equilibrium toward the desired ester products, a stoichiometric excess of alcohol is employed, typically a 6:1 molar ratio of alcohol to ester (or 6:1 methanol to triglyceride for oils, exceeding the theoretical 3:1), which accelerates the reaction and achieves conversions over 95% under optimized conditions. At industrial scales, transesterification can be performed in batch or continuous modes, with batch processes suitable for smaller operations but limited by longer cycle times and inconsistent mixing. Continuous processes, using tubular reactors or static mixers, offer higher throughput but face challenges with viscous feedstocks like unrefined oils, where poor dispersion leads to mass transfer limitations and reduced yields unless enhanced by high-shear mixing or co-solvents. Safety considerations are paramount due to the reactive nature of reagents. Acid-catalyzed reactions, often using sulfuric or , are highly corrosive and necessitate equipment constructed from or corrosion-resistant linings to prevent degradation and contamination. In base-catalyzed systems, free fatty acids or water impurities react to form soaps, which emulsify products and hinder separation, requiring feedstock pretreatment to maintain process integrity.

Mechanism

General Mechanism

Transesterification follows a general addition-elimination common to reactions of esters, proceeding through a tetrahedral regardless of the specific used. This process involves the of the alkoxy group (–OR') of an ester (R–C(=O)–OR') with that of an alcohol (R''–OH), yielding a new ester (R–C(=O)–OR'') and the original alcohol (R'–OH). The reaction is inherently reversible, with the forward and reverse pathways sharing identical mechanistic steps, leading to an equilibrium that favors product formation under conditions of excess alcohol or glycerol removal. The mechanism begins with the nucleophilic attack on the electrophilic carbonyl carbon of the by an ion (RO⁻, derived from the ) or an activated molecule, forming a tetrahedral . In this , the carbonyl carbon adopts sp³ hybridization and is bonded to the original R group, the attacking nucleophilic oxygen (from RO⁻ or R''OH), the original alkoxy group (–OR'), and the former carbonyl oxygen bearing a negative charge. This anionic is stabilized by , where the negative charge delocalizes between the oxygen atoms, potentially involving the alkoxy groups. A representation typically illustrates this step with curved arrows: one depicting the nucleophile's attacking the carbonyl carbon, and another showing the π electrons of the C=O moving to the oxygen, while subsequent structures highlight charge distribution. In the second step, the tetrahedral intermediate undergoes elimination of the (the original , R'O⁻), which collapses the structure and reforms the carbonyl π bond, producing the transesterified product and regenerating the . The curved arrows in diagrams for this elimination phase show the breaking of the C–OR' bond and the reformation of the C=O bond, with the departing as an anion. This step mirrors the reverse reaction, underscoring the nature of the process. A common side reaction occurs in the presence of , where H₂O acts as the instead, leading to and formation of a (R–COOH) and , which can compete with or contaminate the desired transesterification.

Catalyst-Specific Variations

In base-catalyzed transesterification, the catalyst, typically an such as NaOH or KOH, deprotonates the to generate an ion (RO⁻), which serves as a strong that directly attacks the carbonyl carbon of the . This step is represented by the equation: \ce{ROH + B^- -> RO^- + BH} where ROH is the and B⁻ is the base, producing the alkoxide RO⁻ and its conjugate BH. The resulting nucleophilic attack leads to a tetrahedral and proceeds rapidly under mild conditions, but the is equilibrium-limited, often requiring excess to drive yields forward. Acid-catalyzed transesterification initiates with of the carbonyl oxygen in the (RCOOR'), forming an activated species such as R-C(OH)⁺OR', which enhances electrophilicity and facilitates nucleophilic attack by the . This is followed by formation of a tetrahedral and multiple proton transfer steps to eliminate the original and form the new , resulting in a slower overall rate compared to base catalysis but with greater tolerance for and free fatty acids in the feedstock. Enzymatic transesterification, primarily mediated by lipases such as those from , involves the enzyme's binding both the substrate and , where a serine residue in the acts as a to form a covalent acyl-enzyme intermediate. This ping-pong bi-bi allows for sequential acyl transfer, enabling high , such as preferential action at the 1,3-positions of triglycerides, which minimizes unwanted side products. Base catalysis excels in high-throughput applications due to its rapid , acid catalysis suits recalcitrant esters with high free content where methods falter, and enzymatic approaches provide superior specificity for substrates like oils. However, catalysts are susceptible to poisoning by free acids, which promote and reduce efficiency, while acid catalysts often lead to equipment owing to their strong protonating nature.

Types

Alcoholysis

Alcoholysis represents a specific subtype of transesterification in which an (R''OH) acts as the , displacing the alkoxy group from an to form a new and a different as the byproduct; this process is the most prevalent method for modifying structures in . A classic example of alcoholysis involves the of with in the presence of an acid catalyst, yielding and according to the equation CH₃COOCH₃ + CH₃CH₂OH ⇌ CH₃COOCH₂CH₃ + CH₃OH. This exemplifies the exchange of alkyl groups between the ester and the incoming alcohol. The kinetics of alcoholysis typically follow a second-order rate law, with the overall rate depending on the concentrations of the ester and alcohol. The reaction rate is influenced by the chain length and branching of the alcohol, where longer or branched chains introduce steric hindrance that suppresses conversion. One key advantage of alcoholysis lies in the production of a clean alcohol byproduct, which can be readily separated by distillation to shift the reversible equilibrium toward the desired ester product, enhancing yield without introducing complex waste streams. In fragrance and flavor synthesis, alcoholysis is employed to produce esters such as cis-3-hexenyl acetate from corresponding alcohols and acyl donors, leveraging enzymatic catalysts for selective formation of aroma compounds. On a larger scale, alcoholysis of vegetable oils with serves as the basis for , though detailed applications are covered elsewhere. The mechanism proceeds via a tetrahedral , consistent with pathways.

Acidolysis and Thiolysis

Acidolysis refers to the exchange reaction between an ester and a carboxylic acid, resulting in a new ester and a different carboxylic acid, typically represented as RCOOR' + R''COOH → RCOOR'' + R'COOH. This variant of transesterification is thermodynamically challenging due to the similar acid strengths of the reacting and product carboxylic acids, leading to an equilibrium constant near unity that favors incomplete conversion. To drive the reaction forward, an excess of the carboxylic acid is often required, along with elevated temperatures or catalysts such as strong acids or enzymes. A representative example involves the acidolysis of poly(ethylene terephthalate) waste with succinic acid, yielding bis(2-hydroxyethyl) succinate and terephthalic acid under solvent-free conditions at 180–220°C, highlighting its utility in polymer recycling despite side reactions like decarboxylation. Industrially, acidolysis remains rare owing to these equilibrium limitations and competing hydrolysis or dehydration pathways, though it finds niche applications in specialized acyl group transfers. Thiolysis, in contrast, involves the reaction of an with a to form a and an , as in RCOOR' + R''SH → RCOSR'' + R'OH. This process proceeds more rapidly than alcoholysis because thiols are superior nucleophiles compared to s, facilitated by the lower of thiols (around 10–11) which enhances their reactivity under neutral or mildly basic conditions. Transition-metal-free methods enable selective C–O bond cleavage and C–S bond formation at , with broad tolerance for functional groups including aryl halides and alkenes, as demonstrated in the conversion of to S-phenyl thiobenzoate in 85% yield using dimethylphenylsilanol as an additive. In , thiolysis is particularly valuable for generating peptide s, which serve as activated intermediates for native chemical ; for instance, peptide can be converted to alkyl thioesters via NaNO₂ oxidation followed by thiolysis with benzyl mercaptan, achieving yields up to 90% for sequences up to 20 residues. Biochemically, thiolysis plays a key role in acyl transfer reactions, such as those involving , where s facilitate energy-rich intermediates in metabolic pathways like β-oxidation. Unlike alcoholysis, which typically requires harsher conditions for equilibrium shifts due to comparable nucleophilicities, acidolysis and thiolysis are more reversible under standard conditions but can be tuned for selectivity through nucleophile excess or enzymatic , as seen in lipase-mediated variants.

Interesterification

Interesterification refers to the of acyl groups between two molecules or within the ester linkages of a , such as triglycerides in fats and oils, without changing the total number of ester bonds. This rearranges the fatty acid chains on the glycerol backbone, modifying the physical and functional properties of the while preserving their . The interesterification process can proceed randomly or in a directed manner to alter characteristics like melting points, which is particularly useful in applications. In random interesterification, fatty acids are redistributed evenly across all positions until is reached, resulting in a near-uniform distribution among possible species. Directed interesterification, in contrast, employs selective catalysts to favor specific positional exchanges, such as at the sn-1 and sn-3 positions of , enabling tailored structures like equivalents. Catalysts for random interesterification typically include (0.05-0.6% w/w), which facilitates the reaction in a phase at temperatures below 100°C for about 30 minutes. Enzymes, such as sn-1,3-specific lipases (e.g., Lipozyme 435), are used for directed processes, allowing milder conditions and greater control over placement. A representative example is the interesterification of triolein (trisunsaturated triglyceride with ) and tristearin (trisaturated triglyceride with ), which yields a mixture of mixed triglycerides, including 1,2-dioleoyl-3-stearoyl-sn-glycerol and others, enhancing plasticity and spreadability in margarines. This shifts the melting profile to intermediate temperatures, improving texture without trans fats. In shortenings, interesterification enhances oxidative by distributing unsaturated fatty acids more evenly, reducing susceptibility to peroxidation and achieving higher oxidative stability index (OSI) values compared to physical blends, with equilibrium reflecting full randomization of acyl groups.

Applications

Biodiesel Production

is primarily produced through the alkali-catalyzed transesterification of derived from vegetable oils or animal fats with , resulting in methyl esters (), the main biodiesel component, and as a coproduct. This process, often using (NaOH) as the catalyst, occurs in three sequential steps: the first reacts with to form a and one of ; the then converts to a and a second ; finally, the yields and the third . The reaction requires a -to-oil of approximately 6:1, with typical conditions of 50-65°C and 1-1.5 hours reaction time to achieve high conversion. Yields of 95-98% are attainable following purification steps such as washing and distillation to remove residual , catalyst, and . Feedstocks with high free (FFA) content, common in waste oils, necessitate a pretreatment via acid-catalyzed pre-esterification to convert FFAs to methyl esters and prevent soap formation during the subsequent base-catalyzed transesterification. This two-stage approach ensures compatibility with low-FFA oils like or , while enabling utilization of cost-effective high-FFA sources such as used . On an industrial scale, employs continuous-flow reactors, such as or continuous stirred-tank configurations, in facilities processing up to 100,000 tons annually for efficient, high-volume output. requires significant input, typically around 5-10 per kg of , primarily from heating and mixing. Global production reached approximately 50 billion liters of in , with estimates around 45-50 billion liters in as of available data. In the , member states implement mandates under the Directive, often requiring blending levels around 7% or more for renewables in , supporting market growth and integration into transportation . Significant challenges in biodiesel production include managing the glycerol byproduct, which comprises roughly 10% by weight of the output and contains impurities like , water, and soaps that demand costly purification or valorization into chemicals such as . Additionally, the base-catalyzed process exhibits high sensitivity to water, as even trace amounts promote reactions that emulsify the mixture, complicate , and lower FAME yields. These issues underscore the need for dry feedstocks and rigorous to maintain economic viability.

Polyester Synthesis

Transesterification plays a role in the industrial synthesis of polyesters, particularly in one major route for polyethylene terephthalate (PET), through the reaction of a diester such as dimethyl terephthalate (DMT) with a diol like ethylene glycol (EG); however, the predominant method uses direct esterification of purified terephthalic acid (PTA) with EG. In the DMT-based melt-phase process, DMT undergoes transesterification with excess EG to form bis(2-hydroxyethyl) terephthalate (BHET) monomers and release methanol as a byproduct, initiating the formation of ester linkages. The synthesis proceeds in two main stages: an initial transesterification phase at temperatures around 150–250°C, where DMT and EG react to produce low-molecular-weight oligomers and methanol, which is continuously removed to shift the equilibrium forward; this is followed by a polycondensation stage under vacuum at higher temperatures (up to 280°C) to eliminate excess EG and build high-molecular-weight chains (typically Mw > 50,000 Da) through further ester exchanges and condensation. Catalysts such as titanium alkoxides (e.g., tetrabutyl titanate) are commonly employed to accelerate these exchanges, enabling efficient production of PET with the required viscosity for applications like fibers and bottles. Global production exceeds 36 million metric tons annually as of 2023, with major uses in (e.g., beverage bottles) and textiles (e.g., synthetic fibers). The DMT process emphasizes sustainability through methanol recovery, where the distilled is purified and reused, reducing costs and environmental impact.

Fat and Oil Processing

In fat and oil processing, interesterification serves as a key application of transesterification to rearrange the fatty acids within triglycerides of edible oils, thereby modifying physical properties such as , , and spreadability without altering the overall composition or introducing new functional groups. This process is particularly valuable for producing zero-trans fats from sources like , where blends of and vegetable oils (e.g., sunflower or ) are interesterified to create structured suitable for food products. For instance, interesterification of -based blends eliminates the need for partial , which traditionally generates trans fats, while achieving desired textures for applications like and . Two primary methods are employed: chemical interesterification and enzymatic interesterification. Chemical interesterification typically uses (NaOMe) as a catalyst at temperatures below 100°C, often around 90°C, in a batch process lasting about 30 minutes, which randomizes fatty acids across all glycerol positions. In contrast, enzymatic interesterification employs lipases, such as sn-1,3-specific enzymes like those from Rhizomucor miehei (e.g., Lipozyme RM IM), allowing for positional control that preserves fatty acids at the sn-2 position, which is beneficial for digestibility in products mimicking human milk fat. Enzymatic methods operate at milder conditions (around 50-70°C), reduce by-product formation, and enable continuous processing, though they are more costly than chemical approaches. The benefits of interesterification in this context include improved texture and functionality for margarines and shortenings, where it enhances spreadability and creaminess at while maintaining solidity at higher temperatures, often through blending high-melting saturated fats with unsaturated oils to optimize solid fat content profiles. This replaces , avoiding trans fats linked to health concerns, and has been affirmed as (GRAS) by the FDA for enzyme preparations used in the process since 1998. Interesterified fats now hold a significant presence in the , particularly in shortenings and margarines, supporting the production of low-trans or zero-trans products that meet consumer demand for healthier alternatives. Quality control in interesterified fat production relies on () analysis to assess the degree of randomization, typically by examining triacylglycerol (TAG) profiles to confirm even distribution of fatty acids (e.g., targeting 33% at each sn-position for full randomization) and verifying the absence of trans isomers. This ensures product consistency, with solid fat content (SFC) measured via pulsed to evaluate physical performance across temperature ranges.

Organic Synthesis

Transesterification serves as a versatile tool in laboratory for interconversions, particularly in the manipulation of protecting groups and the preparation of reactive intermediates. One key application involves the exchange of protecting groups on alcohols, such as converting an (R-OCOCH₃) to a benzoate (R-OCOC₆H₅), which allows for selective protection strategies in multi-step syntheses without the need for full deprotection and re-protection sequences. This process is often facilitated by base- or enzyme-catalyzed conditions that promote shifts toward the desired product, enabling precise control in complex molecules like carbohydrates or peptides. A prominent example of transesterification in synthesis is the formation of enol ethers, which are valuable building blocks for further transformations such as Diels-Alder reactions or precursors. In this reaction, reacts with an (ROH) to yield the corresponding vinyl ether (RO-CH=CH₂) and acetic acid, driven by the tautomerization of the intermediate to , rendering the process irreversible. The general equation is: \ce{CH2=CH-OC(O)CH3 + ROH -> CH2=CH-OR + CH3COOH} This method is particularly efficient under mild, metal-catalyzed conditions, such as catalysis, accommodating a range of primary and secondary with high yields. Enzymatic transesterification, particularly using lipases, offers exceptional selectivity for kinetic resolution of racemic , achieving enantiomeric excesses () exceeding 99% in many cases. Lipases from sources like Candida antarctica selectively acylate one of a secondary with an activated ester like , leaving the unreacted enriched in optical purity for downstream applications. The development of these enzymatic methods gained momentum in the , enabling their integration into pharmaceutical synthesis, including the preparation of chiral intermediates for statins such as and simvastatin, where high stereocontrol is critical. These approaches provide significant advantages in , operating under mild conditions (often and neutral ) that avoid harsh and minimize side reactions with sensitive functional groups. Moreover, enzymatic variants are scalable to kilogram batches in settings, supporting efficient production of fine chemicals and candidates without extensive purification steps.

History and Developments

Early Discoveries

The earliest documented observation of transesterification dates to 1853, when Irish chemists E. Duffy and J. Patrick conducted experiments on the reaction of vegetable oils with alcohols, demonstrating the exchange of groups to form new and . This work laid the groundwork for understanding the process, though it was initially viewed as a curiosity in rather than a practical method. Subsequent studies in the late , including those exploring exchanges during , further illuminated the reaction's potential, with early hints of tetrahedral intermediates proposed by chemists like Claisen in 1887. In the , transesterification transitioned toward industrial relevance through fat interesterification, particularly for production. Wilhelm Normann, known for his hydrogenation patents, secured a for chemically rearranging fatty acids in triglycerides using catalysts, enabling the conversion of liquid oils into solid fats suitable for soaps and amid shortages of animal fats. This application highlighted the reaction's utility in modifying structures without , marking an early practical adoption in the fats and oils industry. The 1930s and 1940s saw intensified research driven by wartime needs, especially for production to manufacture explosives like during . Efforts to extract from triglycerides via alcoholysis revealed key insights into transesterification kinetics and catalysis, as researchers optimized base-catalyzed processes to yield alkyl esters alongside . A pivotal development was the 1937 Belgian patent by Georges Chavanne for the alcoholysis of s with ethanol or methanol, which separated while producing esters suitable as fuels—often cited as an early precursor. Pre-WWII experiments in further applied these principles, using transesterified to power heavy-duty vehicles, demonstrating the reaction's viability for alternatives. By the 1950s, academic focus shifted to mechanistic elucidation, confirming the role of tetrahedral intermediates in the pathway of transesterification. Studies emphasized the addition-elimination sequence, where attacks the carbonyl, forming a transient tetrahedral before eliminating the original group. Concurrently, industrial advancements included Colgate-Palmolive's patents for continuous transesterification processes, such as U.S. 2,494,366 (1950), which refined recovery from fats and oils, bridging laboratory insights to scalable production.

Modern Advances

Amid rising oil prices following the 1973 embargo, there was renewed interest in via transesterification, building on earlier patents from companies like Colgate-Palmolive-Peet in the . These foundational processes emphasized efficient alcoholysis under basic conditions, laying groundwork for scalable synthesis. Enzymatic transesterification gained significant research traction in the for potential , with pilot and small-scale plants utilizing lipases emerging in the to catalyze from oils and fats, offering milder reaction conditions and reduced byproduct formation compared to chemical . A significant breakthrough came in 2001 with the Japanese supercritical method developed by and Kusdiana, which enables non-catalytic transesterification of triglycerides at 350°C and 20 MPa, achieving up to 99% yield in just minutes without soaps or purification challenges. This catalyst-free approach addresses limitations of traditional methods by simplifying and enhancing efficiency for high-volume production. In parallel, advancements have focused on heterogeneous catalysts like MgO, which provide high activity in transesterification while enabling easy separation and reuse over multiple cycles, minimizing waste and costs in sustainable synthesis. Biocatalysts, particularly immobilized lipases, have been integrated into continuous flow systems, allowing steady-state operation with improved productivity and stability for large-scale applications. The 2010s saw intensified research on feedstocks, such as used cooking oils, for transesterification to produce , leveraging low-cost, abundant resources to reduce environmental impact and economic barriers. In the during the 2020s, policies under the Renewable Energy Directive have promoted advanced biofuels through transesterification pathways, targeting a combined 5.5% for advanced biofuels and of non-biological origin by 2030 to decarbonize transport while ensuring criteria for feedstocks like algal oils and residues. As of 2025, progress includes expanded pilot facilities for enzymatic and supercritical methods, alongside increased glycerol valorization in biorefineries to enhance overall process . Beyond fuels, transesterification via PET methanolysis has emerged as a key technology, depolymerizing waste into monomers like and with over 90% recovery rates, enabling closed-loop production and significant reduction in plastic waste.

References

  1. [1]
  2. [2]
    Transesterification - Master Organic Chemistry
    Nov 10, 2022 · Transesterification is the conversion of one ester into another through replacement of an OR groups, either with an alkoxide or an alcohol ...
  3. [3]
    Recent advances in transesterification for sustainable biodiesel ...
    Jan 23, 2024 · Transesterification in biodiesel production. Alcohol and lipids react chemically to produce fatty acid alkyl esters, a process known as ...
  4. [4]
    [PDF] Determination of Total Lipids as Fatty Acid Methyl Esters (FAME) by ...
    3.3 Transesterification: The process of exchanging the organic group of an ester with the organic group of an alcohol. 3.4 Extractives/Extracted Lipids: The ...
  5. [5]
    Application of corncob residue-derived catalyst in ... - BioResources
    Nov 11, 2019 · Glycerol carbonate can be synthesized by the transesterification of glycerol with dimethyl carbonate (DMC). This method has many advantages, ...
  6. [6]
    Polyester transesterification through reactive blending and its ...
    Jun 6, 2025 · Commonly known for its role in biodiesel production [9], transesterification also plays a significant role in polymer chemistry, such as ring- ...
  7. [7]
    Transesterification - an overview | ScienceDirect Topics
    Transesterification is an imperative process for biodiesel production, as it can reduce the viscosity of the feedstock/vegetable oils to a level closer to the ...
  8. [8]
    Illustrated Glossary of Organic Chemistry - Transesterification
    The reaction is a reversible equilibrium, so methanol is used in excess (it is the reaction solvent) to push the reaction equilibrium towards ethyl benzoate ( ...
  9. [9]
    Transesterification - an overview | ScienceDirect Topics
    All three types are reversible, so that the equilibrium has to be shifted to the desired side either by applying an excess of the reagent or by removing a ...
  10. [10]
    Kinetics of the NaOH-catalyzed transesterification of sunflower oil ...
    The dependency of the equilibrium constants (KC(T)) on temperature was expressed according to the integrated Van't Hoff equation taking 50 °C (323 K) as one of ...
  11. [11]
    Transesterification and esterification for biodiesel production
    To enhance biodiesel production, an excess of alcohol is often used, in accordance with Le Chatelier's principle, to shift the equilibrium toward the desired ...
  12. [12]
    Transesterification - Chemistry Steps
    Transesterification is the conversion of one ester to another – simple: Just like the Fischer esterification and ester hydrolysis, it is also a reversible ...
  13. [13]
    Biodiesel Production Using Homogeneous, Heterogeneous, and ...
    Sep 28, 2021 · Transesterification/esterification is a chemical conversion of lipids such as vegetable oils and animal fats reacting with alcohol to form fatty ...
  14. [14]
    State of the Art of Catalysts for Biodiesel Production - Frontiers
    Heterogeneous catalysts can be divided into acid and base catalysts. These catalysts can be classified as Brønsted or Lewis catalysts (Di Serio et al., 2008).<|control11|><|separator|>
  15. [15]
    Modeling the Effects of Temperature in Enzymatic Biodiesel ...
    Sep 18, 2024 · They have observed the effect of reaction parameters such as reaction time (3–7 h), temperature (35–60 °C), enzyme loading (0.1–0.3 g), and ...
  16. [16]
    Preparation of Ester Derivatives of Fatty Acids for Chromatographic ...
    As with other transesterification procedures, it is usually advisable to add an inert solvent such as toluene to solubilize triacylglycerols. There is a ...Acid-Catalysed... · Diazomethane And Related... · Pyrolysis Of...
  17. [17]
    Impact of Various Catalysts on Transesterification of Used Cooking ...
    Nov 5, 2022 · The presence of water and free fatty acid leads to soap formation via the hydrolysis of triglycerides, which can reduce the yield of the ...Missing: stainless steel
  18. [18]
    [PDF] Basic Chemistry of Biodiesel - ATTRA – Sustainable Agriculture
    We typically use double the stoichiometric ratio, or 6:1 moles of alcohol to triglyceride, to push the reaction to the products side. Page 20. Base Catalysts ...
  19. [19]
    Biodisel Transesterification: Chemistry, Catalysts & Process
    Sep 29, 2025 · Base-Catalyzed Transesterification​​ Challenges involve sensitivity to water and free fatty acids, which lead to soap formation and yield losses. ...<|control11|><|separator|>
  20. [20]
    [PDF] A critical review on production of biodiesel from various feedstocks
    Apr 27, 2016 · Transesterification mechanism, the carbonyl carbon of the starting ester (RCOOR1) undergoes nucleophilic attack by the incoming alkoxide ...
  21. [21]
    [PDF] Kinetics of Transesterification Processes for Biodiesel Production
    Then, the tetrahedral carbon is separated from the intermediate to form an alkyl ester (third step). Deprotonation of catalyst regenerates the alkali, whereas ...
  22. [22]
    [PDF] Transesterification - L.S.College, Muzaffarpur
    Aug 22, 2020 · In the transesterification mechanism, the carbonyl carbon of the starting ester reacts to give a tetrahedral intermediate, which either reverts ...
  23. [23]
    A Density Functional Theory Study on the Acid‐Catalyzed ...
    Dec 14, 2018 · The acid-catalyzed transesterification reaction has been one of the most effective methods to produce biodiesel, especially from waste ...
  24. [24]
    Lipase-catalysed transesterification: Viewpoint of the mechanism ...
    This paper reviews the current application of MOFs ... lipase resources, intensification technics, and process modelling for enzymatic transesterification.
  25. [25]
    Comparison of Chemical and Enzymatic Methods for the ...
    Jan 17, 2020 · In the transesterification reactions catalyzed by immobilized enzymes, the thermal stability of the catalysts required the reaction temperature ...
  26. [26]
    An encapsulated report on enzyme-assisted transesterification with ...
    This report summarizes the various sources of lipase, various production strategies for lipase and the lipase-mediated transesterification.
  27. [27]
    Esters in the Food and Cosmetic Industries: An Overview of the ... - NIH
    Esters can be synthesized by means of esterification between acids and alcohols, as well as through transesterification, alcoholysis, or acidolysis reactions.Missing: examples | Show results with:examples
  28. [28]
    Kinetic study of transesterification of methyl acetate with ethanol ...
    PDF | The chemical equilibrium and reaction kinetics for transesterification of methyl acetate with ethanol to form ethyl acetate and methanol catalyzed.
  29. [29]
    [PDF] Kinetics of Transesterification - NC State Repository
    Alkyl esters are commonly made from vegetable oil through a chemical reaction called transesterification. Although this process has had long use in making ...
  30. [30]
    Kinetics of Dowex 50W catalyzed transesterification of methyl ...
    Feb 6, 2024 · It was found that an increase in the chain length of the alcohol and/or branching suppressed the conversion of the reactants owing to steric ...Missing: second- | Show results with:second-
  31. [31]
    [PDF] Methyl acetate transesterification by reactive pressure-swing ...
    Feb 2, 2018 · The alcoholysis of methyl acetate to obtain more valuable products from the PVA process residual stream has been thoroughly studied when it ...
  32. [32]
    Current developments in esterification reaction: A review on process ...
    The mass transfer can be enhanced by reducing the film thickness and shifting the equilibrium in forward direction by simultaneous removal of a byproduct.
  33. [33]
    Solvent-Free Synthesis of Flavour Esters through Immobilized ... - NIH
    The effect of different transesterification variables, namely, molarity of alcohol, reaction time, temperature, agitation, addition of water, and enzyme amount ...
  34. [34]
    Natural flavor ester synthesis catalyzed by lipases - Bayout - 2020
    Nov 15, 2019 · The synthesis of three flavour esters, namely cis-3-hexenyl acetate, via transesterification and esterification of the natural corresponding ...
  35. [35]
    [PDF] Investigations of Cost-Effective Biodiesel Production from High FFA ...
    Transesterification or alcoholysis is a reversible chemical reaction of a vegetable oil, animal fat, or algal oil (mainly composed by triglycerides) with an ...
  36. [36]
    Acidolysis of Poly(ethylene terephthalate) Waste Using Succinic ...
    Dec 8, 2022 · The exchange reaction of small molecules between an ester group and a carboxylic acid, called carboxylic acidolysis or briefly acidolysis, is ...
  37. [37]
    Insight into Chemical Recycling of Flexible Polyurethane Foams by ...
    Jan 11, 2022 · Acidolysis is emerging as a promising method for recycling polyurethane foam (PUF) waste. Here, we present highly efficient acidolysis of PUFs with adipic acid ...
  38. [38]
    Conversion of esters to thioesters under mild conditions
    We report conversion of esters to thioesters via selective C–O bond cleavage/weak C–S bond formation under transition-metal-free conditions.
  39. [39]
    Peptide Alkyl Thioester Synthesis from Advanced Thiols and Peptide ...
    Peptide alkyl thioesters are versatile reagents in various synthetic applications, commonly generated from peptide hydrazides and thiols.Experimental Section · Supporting Information · Author Information · References
  40. [40]
    Thioesters provide a plausible prebiotic path to proto-peptides - Nature
    May 11, 2022 · It is known that thiols can condense with carboxylic acids to form thioesters (Fig. 1, top). Further, thioester linkages can be converted to ...
  41. [41]
    Biodiesel Synthesis by Enzymatic Transesterification of Palm Oil ...
    This work deals with the transesterification of palm oil with ethanol in a solvent free system using lipase from different sources.
  42. [42]
    Chemical Interesterification - AOCS
    Interesterification can be defined as a redistribution of the fatty acid moieties present in a triglyceride oil over its glycerol moieties.
  43. [43]
    Enzymatic Interesterification - AOCS
    Enzymatic interesterification is a process in which a lipase enzyme is used to catalyse the exchange of fatty acids attached to the glycerol backbone of a fat.
  44. [44]
    Interesterified fats: What are they and why are they used? A briefing ...
    Oct 23, 2019 · Interesterification is a means of modifying the structure and functionality of fats and oils to produce food ingredients for a range of ...
  45. [45]
    Enzymatically Interesterified Triadica sebifera Oil: A Novel ... - MDPI
    Additionally, the shortening demonstrated better oxidative stability (OSI) than the physical blends, ensuring a longer shelf life. Beyond its functional and ...Missing: randomization | Show results with:randomization
  46. [46]
    8.2 The Reaction of Biodiesel: Transesterification | EGEE 439
    The reaction is called transesterification, and the process takes place in four steps. The first step is to mix the alcohol for reaction with the catalyst, ...
  47. [47]
    Biodiesel Production from Alkali-Catalyzed Transesterification ... - NIH
    May 18, 2022 · Biodiesel production using transesterification is an economical and time-saving method and produces fuel with higher cetane number and improved ...
  48. [48]
    Pre-esterification of high acidity animal fats to produce biodiesel
    Pre-esterification of these raw materials is an effective pre-treatment to reduce FFA levels as free fatty acids react with methanol to produce methyl esters ...
  49. [49]
    [PDF] Esterification Pretreatment of Free Fatty Acid in Biodiesel Production ...
    In the US, biodiesel producers usually follow the 19.8:1 methanol-to-FFA molar ratio for free fatty acid (FFA) es- terification, as suggested by the ...
  50. [50]
    The scale up biodiesel production reactor pilot plant and industrial ...
    Jan 29, 2025 · Meanwhile, on an industrial scale, the biodiesel yield for CSTR reactor was 79,7% and PFR was 87,3% at production capacity of 100.000 tons/year.
  51. [51]
    Energy and cost analyses of biodiesel production from waste ...
    The total energy input and energy output were calculated as 30.05 and 44.91 MJ L−1, respectively. The energy output/input ratio was 1.49 in biodiesel production ...
  52. [52]
    Biofuels explained - data and statistics - U.S. Energy ... - EIA
    Biodiesel data ; U.S. imports, 2022, 250 ; U.S. exports, 2022, 238 ; U.S. consumption, 2022, 1,658 ; World production, 2021, 13,966 ; World consumption, 2021, 13,776 ...
  53. [53]
    [PDF] Report Name:Biofuel Mandates in the EU by Member State - 2024
    Jun 27, 2024 · Per the current legislation, the maximum blend with conventional fuel cannot exceed seven percent for biodiesel, and ten percent for bioethanol, ...
  54. [54]
    Environmental impacts of valorisation of crude glycerol from ...
    Apr 30, 2024 · Biodiesel production generates about 10 % w/w crude glycerol as a by-product, which contains various impurities, including water, methanol, soap ...
  55. [55]
    An overview to process design, simulation and sustainability ...
    Jun 1, 2021 · A process model to estimate the cost of industrial scale biodiesel production from waste cooking oil by supercritical transesterification.
  56. [56]
    Improved polymerization and depolymerization kinetics of poly ...
    Jul 5, 2021 · Industrially, PET is synthesized either by transesterification of dimethyl terephthalate (DMT) with ethylene glycol (EG) (DMT process) or ...
  57. [57]
    Industrial Chemistry Module | English - University of Scranton
    This is a transesterification reaction in which one ester is transformed into another.The synthesis of PET by this method reacts dimethylterephthalate with ...
  58. [58]
    PET synthesis reactions (a) esterification of TPA with EG (b)...
    The second reaction is a trans-esterification reaction (Figure 2b) where dimethyl terephthalate (DMT) reacts with EG at 180 -210°C and 100 kPa [3]. Trans- ...
  59. [59]
    Mechanistic Details of the Titanium-Mediated Polycondensation ...
    In this work, the mechanism of polyester polycondensation catalysed by titanium catalysts was investigated using density functional theory (DFT).
  60. [60]
    [PDF] Kinetic and Catalytic Studies of Polyethylene Terephthalate Synthesis
    Mar 1, 2010 · Table 2.1 names the companies which produce PET with titanium catalysts. However, these PET products are mostly used in fiber application. Table ...
  61. [61]
    Polyethylene Terephthalate Market Size, Share & Forecast 2035
    The global PET market was around 30 million tonnes in 2024 and is expected to reach 44 million tonnes by 2035, with a 3.81% CAGR. Food & Beverage Packaging ...
  62. [62]
    Production of Trans‐free fats by chemical interesterified blends of ...
    Oct 3, 2019 · In this study, production of trans-free fats through chemical interesterification of binary blends of palm stearin (PS) and sunflower oil (SFO)
  63. [63]
    (PDF) Interesterified palm oil as alternatives to hydrogenation
    Aug 9, 2025 · There is no formation of trans fatty acids upon interesterification. Interesterified fats are widely used as textural fats for solid fat ...
  64. [64]
    Enzymatic interesterification effect on the physicochemical and ...
    Blends solid fat content behavior was affected by the interesterification process, mainly in the temperature range of 10–20 °C, reducing the solids content by ...Missing: NaOMe | Show results with:NaOMe
  65. [65]
    What are interesterified fats and should we be worried about them in ...
    Interesterified (IE) fats are now used as a hard fat replacement for trans fats in a large range of commonly consumed foods, including spreads, bakery and ...
  66. [66]
    A GC/MS method for the rapid determination of disaturated ...
    May 30, 2023 · Fats chemically randomized or produced by chemical interesterification have a random distribution of saturated fatty acids in their TAG ...
  67. [67]
    Protection (and Deprotection) of Functional Groups in Organic ...
    In this review we describe the application of heterogeneous catalysis in protecting group chemistry, focusing mainly on that developed during the past decade.
  68. [68]
    Protecting group migrations in carbohydrate chemistry - ScienceDirect
    Protecting groups are valuable in chemo- and regioselective synthetic manipulations. In particular, they are indispensable in carbohydrate chemistry.Missing: benzoate | Show results with:benzoate
  69. [69]
    Use of Iridium‐Catalyzed Transfer Vinylation for the Synthesis of Bio ...
    Jan 19, 2022 · The iridium catalyzed transfer vinylation of bio-based polyols and of other alcohols and phenols with interesting structural motifs was accomplished
  70. [70]
    Enzymatic Kinetic Resolution and Chemoenzymatic Dynamic Kinetic ...
    The combination of the enzymatic kinetic resolution with a ruthenium-catalyzed alcohol racemization led to an efficient dynamic kinetic resolution (ee up to 99% ...
  71. [71]
    Biocatalyzed Synthesis of Statins: A Sustainable Strategy for ... - MDPI
    In this review, we will comment on the pleiotropic effects of statins and will illustrate some biotransformations nowadays implemented for statin synthesis.
  72. [72]
    Ester synthesis by transesterification - Organic Chemistry Portal
    Transesterification/acylation reactions of secondary alcohols are efficiently catalyzed by N-Heterocyclic carbenes (NHC) at room temperature. R. Singh, R. M. ...
  73. [73]
    Power of Biocatalysis for Organic Synthesis | ACS Central Science
    Jan 14, 2021 · Biocatalysis, using defined enzymes for organic transformations, has become a common tool in organic synthesis, which is also frequently applied in industry.
  74. [74]
    Biodiesel from vegetable oils via transesterification in supercritical ...
    Patrick conducted it as early as 1853. One of the first uses of transesterified vegetable oil was powering heavy-duty vehicles in South Africa before World War ...
  75. [75]
    Tetrahedral Intermediates in Reactions of Carboxylic Acid ...
    Aug 6, 2025 · Already in 1887, Claisen suggested a tetrahedral intermediate in transformations of carboxylic acid derivatives with nucleophiles. A historical ...
  76. [76]
    Trends in Fat Modifications Enabling Alternative Partially ... - NIH
    Jun 1, 2023 · Hence, enzymatic interesterification has found applications in the fat industry for producing specialty functional fats with specific TAG ...
  77. [77]
    [PDF] 80 years of Biodiesel Georges Chavanne
    May 11, 2017 · • The patent describes the alcoholysis (transesterification) of vegetables using ethanol or methanol in order to separate the glycerol from ...
  78. [78]
  79. [79]
    [PDF] Biodiesel Production Technology: August 2002--January 2004 - NREL
    The catalyst used for carrying out the transesterification is usually sodium hydroxide (NaOH) or potassium hydroxide (KOH). These compounds belong to a class of ...
  80. [80]
    Moving towards a Competitive Fully Enzymatic Biodiesel Process
    In the 1990s, the first industrial enzymatic ... Combining enzymatic esterification with conventional alkaline transesterification in an integrated biodiesel ...
  81. [81]
    Metal oxide-based heterogeneous catalysts for biodiesel production
    It exhibits excellent reusability with no discernible decrease in the catalytic performance after five cycles. MgO synthesized via a sol-gel technique showed a ...
  82. [82]
    Continuous flow reactor based with an immobilized biocatalyst for ...
    The conversion of transesterification reaction products (ethyl esters) was determined using gas chromatography and the quantities of intermediate products ( ...
  83. [83]
    Biodiesel production from waste cooking oil - ScienceDirect.com
    Transesterification of waste cooking oil. Most commonly used method for the synthesis of biodiesel from waste cooking oil is by transesterification reaction.
  84. [84]
    [PDF] ADVANCED BIOFUELS IN THE EUROPEAN UNION
    Further processing options include either transesterification to produce biodiesel or hydrotreating the oils to produce renewable diesel. Algae can be used for ...
  85. [85]
    Polyethylene Terephthalate (PET) Recycled by Catalytic Glycolysis
    Efficient recovery of EG and methanol. -. High final product quality. -. Easy purification steps. -. Severe reaction conditions. -. High costs of monomer ...<|control11|><|separator|>