Next Article in Journal
Theoretical Study of Ni- and Cu-Doped Molybdenum Ditelluride Electrocatalysts for Carbon Dioxide Reduction to Formic Acid and Carbon Monoxide
Previous Article in Journal
Carbon-Coated Magnetic Catalysts for Enhanced Degradation of Nitrophenols: Stability and Efficiency in Catalytic Wet Peroxide Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Immobilized Lipases in the Synthesis of Short-Chain Esters: An Overview of Constraints and Perspectives

by
Ronaldo Rodrigues de Sousa
1,2,
Michelle M. dos Santos
1,2,
Matheus W. R. Medeiros
1,
Evelin A. Manoel
3,
Ángel Berenguer-Murcia
4,
Denise Maria Guimarães Freire
2,
Roberto Fernandez-Lafuente
5,* and
Viridiana Santana Ferreira-Leitão
1,2,*
1
Laboratório de Biocatálise, Bioprocessos e Bioprodutos (LABIC) Instituto Nacional de Tecnologia (INT/MCTI), Rio de Janeiro 200081312, RJ, Brazil
2
Programa de Pós-Graduação em Bioquímica/Instituto de Química (PPGBq/IQ), Universidade Federal do Rio de Janeiro—UFRJ, Avenida Athos da Silveira Ramos, 149, Rio de Janeiro 21941-909, RJ, Brazil
3
Departamento de Biotecnologia Farmacêutica, Faculdade de Farmácia, Centro de Ciências da Saúde (CCS), Universidade Federal do Rio de Janeiro—UFRJ, Avenida Carlos Chagas Filho, 373, Bloco K, Rio de Janeiro 21941-972, RJ, Brazil
4
Departamento de Química Inorgánica e Instituto Universitario de Materiales, Universidad de Alicante, 03690 Alicante, Spain
5
Departamento de Biocatálisis, ICP-CSIC, Campus UAM-CSIC, 28049 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(4), 375; https://doi.org/10.3390/catal15040375
Submission received: 11 March 2025 / Revised: 7 April 2025 / Accepted: 9 April 2025 / Published: 11 April 2025
(This article belongs to the Section Biocatalysis)

Abstract

:
Biocatalysis—specifically the use of immobilized lipases—has been proposed as a greener alternative for ester production. Several critical challenges, such as the high cost of biocatalysts, are delaying the industrial implementation of biocatalysis. Moreover, for short-chain ester synthesis, the strong inhibition/inactivation potential of short-chain acids and alcohols on lipases leads to long reaction cycles and/or the need to use organic solvents to overcome the limitations of solvent-free systems and, consequently, the decrease in product concentrations. This review presents an overview of the scientific developments in enzymatic short-chain ester synthesis, compiling the constraints on their syntheses from a process perspective, including insights about key performance indicators (KPI) and economic parameters.

1. Introduction

Ester synthesis has been intensively explored in the scientific literature throughout the past decades [1,2,3,4,5,6]. Biocatalytic routes have gained great relevance in this field due to their environmental benefits toward greener and sustainable processes in chemical industries [7,8,9,10,11]. The adoption of biocatalysis in ester syntheses using lipases (E.C. 3.1.1.3) in immobilized form is challenging considering the high costs involved in the process and other technical drawbacks related to their catalytic activity and/or operational stability [5,8,12]. These drawbacks are more significant when considering the synthesis of short-chain esters, due to the acidity induced by short-chain acids, their relative hydrophilic characteristics, and the enzyme inactivation potential of short-chain alcohols [13,14,15,16,17,18]. Short-chain esters, whose chain length range comprises molecules containing 2 to 10 total carbons, are well known for their high volatility and flavor/fragrance properties. These compounds are derived from short-chain carboxylic acids and alcohols usually found in the metabolism of living organisms. Accordingly, some short-chain esters are naturally occurring compounds that can be extracted from plants and microorganisms [1,3,19,20,21,22,23] but with high costs and low yields. Short-chain esters, such as methyl, ethyl, and butyl acetate, are also important solvents and plasticizers [2,21,23,24,25]; the esters of dicarboxylic acids such as acrylic and succinic acids are precursors of polymers like polybutylsuccinate and polymethylmethacrylate [2,23,26]. Acetates, propionates, butyrates, and valerates are useful as chemical building blocks to obtain more complex molecules like fragrances, in which chemical modifications are required to improve stability or compatibility in personal care product formulations or even polymers [2,21,23,27]. Short-chain esters are stable liquids at room temperature. Most of them show a relatively hydrophilic behavior (logP < 2.0), have moderate solubility in aqueous buffers, are miscible with hydrophilic/moderately hydrophobic solvents, and are generally less dense than water [28,29]. Most of them are non-toxic and biodegradable compounds, which are important features for food ingredients, cosmetics, and personal care products [27,30,31,32,33]. Industrially, due to the high cost of their extraction from natural resources, esters are produced via esterification between a short-chain carboxylic acid and a short-chain alcohol in the presence of a catalyst via nucleophilic attack, forming water as a by-product [2,16,23].
Esterification is a thermodynamically controlled process, which means that yields are marked exclusively by the thermodynamic constant of the reaction; the catalysts can only alter the reaction rates, and yields can be under the thermodynamic value if the biocatalyst suffers inhibition or inactivation, as opposed to transesterification reactions [16,34,35]. The use of an excess of one of the reagents (generally the alcohol) is a possibility to improve the yields regarding the minority substrate; another option is the removal of water (the reaction by-product) to shift the thermodynamic equilibrium toward the desired product [16,17,36,37,38,39,40,41,42,43,44]. Due to the lower volatility of the substrate alcohol, selective distillation of the ester may be complex, and extracting the water by evaporation is not straightforward, as the substrates and products may have a lower evaporation temperature [45,46,47,48,49].
In practical terms, however, both strategies are simultaneously adopted—the use of an excess of alcohol and the removal of the formed water (via adsorption on molecular sieves, evaporation whenever possible, etc.) [45,46,47,50,51,52,53,54]. Short-chain alcohol and short-chain acid molecules frequently have mutual solubility, simplifying the use of solvent-free systems [17,37,43,55,56,57]. In this case, the hydrophobicity of the media will increase along with the reaction progress (as the ester will be more hydrophobic than the acid and the alcohol) with possible partitioning of the most hydrophobic unreacted reagents from the formed water where the most hydrophilic reagents may be concentrated if a water phase is formed [17,40,43,58,59]. If the increase in the hydrophobicity of the reaction medium is high enough when the ester is produced, this partitioning may form an aqueous phase in the system at some point in the reaction course (Figure 1). As the biocatalyst particles will be initially filled with the acid and the alcohol, it is likely that the hydrophobic drops of the ester may be unable to penetrate inside the biocatalyst by capillary forces, forming a kind of biphasic system and enabling higher yields of possible esters, as the product will be extracted from the enzyme environment [58,60,61]. This may be similar to the use of standard biphasic systems to shift the equilibrium in thermodynamically controlled reactions [62,63,64,65] with the difference that this is an in situ formed biphasic system and the researcher has scarce control over the hydrophobic phase, that will ultimately be the reaction product. In the following sections, we will discuss the short-chain esterification reactions in the industrial context and perspectives for biocatalysis in this field.

2. Discussion

2.1. Industrial Relevance and Market Size

Accurate data about production volume and specific market sizes are scarce. Some market research companies claim to offer such estimations [17,66]. For instance, one of the main players in this market, IFF®—International Flavors and Fragrances—which has many short-chain esters in its portfolio, reported sales of USD 11.7 billion in 2021 [67]. The demand for esters as crucial ingredients, functioning as emulsifiers or stabilizers, coupled with the ingredient biodegradability and environmentally friendly characteristics are expected to drive and boost esters demand in many industries [1,68]. The high interest rates applied to innovative ingredients and sustainable processes in some sectors can be attested in the fragrance/cosmetics sector, which promotes annual awards for new products and specific categories for these topics, such as “In-Cosmetics Global Awards”, “Pure Beauty Global Awards”, and “Barcelona Perfumery Congress” [69,70,71].
Considering the range of direct applications in dairy products, it is intuitive to think that high volumes of short-chain esters are produced per year. In the United States alone, the esters market was valued at USD 3.8 billion in 2019 and is projected to surge to USD 5 billion by 2025. In a worldwide scenario, the esters market size was projected to be worth USD 94.19 billion in 2023, and there are prospects that the global market for esters will reach a staggering USD 159.36 billion by 2033 [72,73]. A significant concentration of major industry players is observed in the United States (U.S.), highlighting the influential role these companies (will) play in shaping global trends, innovations, and standards within the esters market [73].
This aforementioned growth is chiefly propelled by the rising need for emulsifiers and stabilizers in the personal care and detergent sectors [1]. Ethyl acetate, which is used as a solvent, chemical building block, or plasticizer, dominates the commercial market due to its biodegradable and environmentally friendly attributes [24,25,30,31,74]. Ethyl acetate, isopropyl acetate, and n-butyl acetate are recommended in the CHEM21 selection guide of solvents, which includes safety, health, and environmental criteria [75]. Some recent news from great worldwide companies like BASF®, Evonik®, and OQ® Chemicals indicate that the production of short esters is continuously growing, as well as the expansion and construction of new facilities dedicated to ester production in recent years [76,77,78,79,80].
The economic viability of producing short-chain esters hinges on the production scale and successful commercialization, given their relatively modest costs [8,66]. Nevertheless, the biocatalytic pathway encounters hurdles in this scenario, primarily attributed to the high costs of commercial biocatalysts, notably immobilized lipases.

2.2. Evolution in the Ester Production Processes Using Standard Catalysts

Generally, industrial production uses conventional catalysts to obtain esters such as strong acids or metallic compounds [1,2,68]. Strong acids (inorganic, such as sulfuric or hydrochloric acid, or organic ones, such as p-toluenesulphonic acid), ion exchange resins, metals, metal oxides, or derivatives are usual catalysts for esterification reactions [2,9,68,81,82,83,84,85,86,87,88]. Acid catalysts act as the proton donor to the carboxylic acid; then, a nucleophilic attack by the alcohol may occur [2,16,37,38]. Jyoti and co-workers (2018) have demonstrated that sulfuric acid is the best catalyst when compared to other inorganic acids or even other heterogeneous catalysts in the syntheses of acrylic esters [85]. Its strong acidity and dehydrating capacity promote a rapid reaction rate. However, sulfuric acid is miscible with many different organic media, leading to additional difficulties in post-reaction operations besides its corrosive characteristics [2,81,85].
Homogeneous catalysts in a liquid–liquid reaction in a batch configuration demands a proper separation between products, non-reacted reagents, and the catalyst [2,68,84,89,90,91]. On the other hand, heterogeneous catalysts enable catalyst recovery by simple filtration or decanting after each reaction cycle [10,21,49,90,91,92,93]. The formation of water is another problem in these systems due to its negative effect on the stability of the catalysts. Thus, water removal is the critical step for obtaining high yields from the thermodynamic and kinetic perspective [17,40,41,49].
Due to the lack of catalytic selectivity and harsh reaction conditions (temperatures close to or higher than 100 °C), careful downstream operations are required for product purification; side products (such as sulphonic acid esters) may bring undesirable color and odor to the obtained ester, and unsaturated carboxylic acids are susceptible to thermal degradation or polymerization [2,68,94,95]. This purification step is more exigent in personal care applications or food/beverage industries, requiring deodorizing and bleaching processes, among other operations [1,68]. The neutralization of waste acid catalysts generally requires large volumes of alkali and costly waste treatment for final disposal [2,9].
To solve these limitations, alternative catalysts have been studied, including copper, tungsten, and zirconium oxides, modified clays, and processed wastes [2,86,88,96,97,98], showing promising results. As sulfuric acid is a widespread low-cost bulk chemical, careful comparison with new catalysts and sulfuric acid is required for future implementation. Additional criteria should be included to shift the chemical technology trend toward greener processes, such as life cycle assessment, other environmental indicators, and labeling/certification of the chemicals as green or “natural” [9,66,99,100,101]. However, during the development of new processes, the economic aspects should be realistic, considering the low percentual contribution of conventional catalysts in operational costs of the current processes [8,12,66,102,103].

2.3. New Trends for Ester Production: Biocatalysis

Cosmetics, personal care, and food products are important markets in which consumers may push the transition of ester production to green routes [1,3,9,68,104]. These industrial sectors are dynamic, demanding frequent changes in formulations and new appeals to the final customers [68]. Enzyme selectivity ensures the lack of undesired side products and their capacity to work under mild conditions, prevent undesired modification of the products, and allows more energy-efficient processes [1,9,68,105]. Esters produced by biocatalysis can be labeled as “green” or “natural” in some specific markets, permitting them to reach higher added values [1,38,68,106,107,108]. Moreover, biocatalytic routes can address many of the principles of Green Chemistry [37,102,109,110] and reduce the carbon footprint and greenhouse gas emissions [9,111,112]. This set of environmental advantages may increase the attractiveness of developing and implementing enzymatic routes in short ester syntheses.
Enzymes are highly selective catalysts that operate under mild reaction conditions; their use in organic syntheses, consequently, leads to the production of purer products, as the selectivity prevents the production of undesired by-products, and the mild conditions prevent decomposition of the reactants [1,9,68,102]. Enzymes used as biocatalysts are considered safe materials, although they may be subjected to some level of legal regulation as preconized in the Regulation (EC) 1907/2006 of the European Parliament concerning the Registration, Evaluation, Authorization and Restriction of Chemicals (REACH) and European Union Classification, Labelling and Packaging Regulation 1272/2008 [113,114]. More specifically, Enzyme TEC 21/19—ERC, describes that the only safety concern is related to a potential respiratory allergy during enzyme handling [115]. The mild reaction conditions under which lipases operate bring forth an energy saving to the process: about 60% less than the conventional process using sulfuric acid [9,68]. Their enantioselectivity permits obtaining a unique enantiomer from a prochiral substrate. Enzymes are also specific catalysts useful for the resolution of racemic mixtures [5,111]. The minimal formation of by-products in an enzymatic reaction facilitates the purification steps of products and increases their quality [1,68,94,107].
Among the possible enzymes that may be useful as biocatalysts for ester production, lipases stand out. Lipases catalyze in vivo metabolism of glycerides to produce free fatty acids and glycerol, but in vitro they can catalyze a diversity of reactions. Besides esterification, they can be used as catalysts in reactions such as transesterifications [16,116,117,118,119,120,121,122], acidolysis [123,124,125,126], interesterifications [127,128,129,130], or amidations [131,132,133,134,135], together with promiscuous reactions [5,136,137]. One of the most remarkable features of lipases is the presence of a mobile polypeptide chain (lid or flat) [138,139,140], that allows two different lipase conformations. The internal face of the lid is quite hydrophobic and interacts with the hydrophobic surroundings of the active center in homogeneous aqueous media (close and usually inactive form). When it moves, a large hydrophobic pocket is exposed to the medium and the active center is also exposed to the medium (open and active form) [138,139,140]. In most instances, the lid fully covers the active site in the closed form. There are cases with some exceptions such as the lipase B from Candida antarctica [140]. The huge hydrophobic pocket of the lipase open form can interact with any hydrophobic surface (like drops of the substrate, such as triglycerides), fixing the open form of the lipase to the drops of the substrate. This phenomenon is known as interfacial activation [138,141,142,143,144], since when the substrate concentration exceeds the solubility concentration and drops of the substrate are formed, a leap in enzyme activity is detected [138,141,142,143,144]. However, this tendency of the lipases to become adsorbed on hydrophobic surfaces may be expanded to any other hydrophobic surface different from oil drops [145], such as gas bubbles [146], some hydrophobic proteins [147,148,149], the open form of another lipase molecule [150,151,152,153], or, as it will be discussed later, a hydrophobic matrix [143].
In this aforementioned context, immobilized lipases have been extensively explored in the past decades for esterification reactions [6,66,111,154]. Despite the potential advantages, the ability of immobilized lipases to catalyze esterification reactions has been neglected in large-scale contexts considering that few industrial companies are applying this route [1,10,11,66,68]. As far as we know, none of them are related to short-chain esters. The main reason is due to technical and economic constraints, particularly the higher costs of immobilized lipases (generally between USD 100–1000 kg−1) compared to the costs of soluble enzymes [8,12,105,155]. Purification and immobilization of lipases contribute greatly to the final costs [8,156], and strategies to obtain cheaper biocatalysts are required for the development of this technology, such as immobilization using agro-industrial wastes or the use of dry fermented solids with enzymatic activity [10,154,157,158]. According to Tüvfesson and co-authors (2011) [8], the maximum allowable cost contribution for the biocatalyst should be EUR 0.05–0.25 kg−1 and productivities should be 400–2000 kg product per one kg of biocatalyst and 2000–10,000 kg product per one kg of biocatalyst, respectively, for production of specialties and commodities. Elevated costs could be compensated if the biocatalysts demonstrate a significant improvement in performance and extended reuse in the reaction [12,159].
The use of high concentrations of acids, which is a requirement to have a high concentration of the ester, may have a negative effect on the lipase performance [13,15,37]. These constraints are usually overcome using hydrophobic solvents, as this reduces the concentration of reagents and the extent of their negative effects [14,36,111,160]. The use of solvents is undesirable in industrial processes for different reasons: (i) reduction of volumetric productivity; (ii) the need for solvent recovery steps; (iii) increase in operational costs, and (iv) potential occupational and environmental risks [14,17,43,144]. Therefore, short-chain ester syntheses using immobilized lipases have an additional challenge—the need for an organic solvent—to achieve feasibility in a large-scale context, apart from the other well-known constraints for biocatalytic routes on ester production. The adoption of green solvents or conventional solvents that allow easy and efficient recycling can be a convenient path when their use is crucially required [161,162]. Another option is the design of lipase biocatalysts able to withstand the presence of high short-chain acids and alcohol concentrations.
Thus, even the operational and environmental advantages of the biocatalytic route—mild reaction conditions, purity of obtained products, and simplification of downstream operations—are overshadowed by the costs of immobilized lipases in the context of obtaining products of low–moderate value. Simple, efficient, and lipase-improving features of lipase immobilization should be developed to enable the use of these biocatalysts for this specific application.

2.4. Immobilization of Enzymes

Enzyme immobilization was initially developed to solve the problem of simplifying the recovery of the initially extremely expensive enzymes, and if the enzyme remains active, its reuse over subsequent reaction cycles [163,164,165]. Moreover, the use of immobilized enzymes allows the user to benefit from the advantages of the application of a heterogeneous catalyst (simplifying reactor control and the downstream of the product) [163,164,165]. Nowadays, the production prices of enzymes, and specifically of lipases, are rapidly decreasing, reducing the necessity for enzyme immobilization. For instance, Eversa® has been launched by Novozymes® as a lipase for biodiesel production to be employed in soluble (non-immobilized) form, and it accounts for only 5% of the production costs [166].
However, researchers tried to couple immobilization, an initially almost compulsory step to use enzymes in industry, to the improvement of other enzyme features [154]. The first goal was to improve enzyme stability, and soon it was shown that if the enzyme was covalently attached to the support via multiple bonds, enzyme rigidity may increase along with enzyme stability [154,167,168]. Later, some other reasons for enzyme stabilization after immobilization were pointed out, as recently reviewed [5,143,169]. However, it should be clear that enzyme stabilization is not a direct consequence of any randomly designed immobilization protocol, but the result of an adequate design of the immobilization process [154]. Immobilization has proved to be a valuable tool in tuning enzyme selectivity, specificity, inhibition, etc., as the enzyme structure is distorted by interacting with the support and this alters their final properties [143,170,171]. Comparative studies between immobilized and soluble lipases in the same synthesis demonstrate that the immobilized enzymes enable to get higher ester yields [172,173,174,175], due to the increase in stability, and specificity, together with the prevention of enzyme aggregation, which avoids the generation of new mass transfer matters [175,176,177]. Moreover, immobilization following a proper protocol can permit it to be coupled to enzyme purification, saving costs and enabling highly mass-activity biocatalysts using crude enzyme preparations [178].
Usual protocols for lipase immobilization include adsorption [143,179,180,181,182], entrapment [183,184,185,186,187,188], encapsulation [189,190,191], covalent coupling [192,193,194], and formation of cross-linked aggregates or crystals [172,173,195,196]. A variety of materials have been used as immobilization matrix–polymers [53,176,178,182,197,198,199,200,201], resins [120,202,203,204,205], inorganic compounds [61,129,205,206], or even waste materials (as fibers, straws, cloth, rice husk ash, and other agricultural waste) [207,208,209,210,211,212]. However, the most popular protocol is lipase immobilization on hydrophobic supports via interfacial activation of lipases against the support surface [143,154]. The hydrophobic supports used for this immobilization strategy are very chemically stable supports (as they have no chemically reactive moieties that can be inactivated or modified during storage or the lipase immobilization process) [143]. This method is very rapid and simple. By just mixing the enzyme solution and the support, the lipase becomes rapidly immobilized [143]. As the lipase adsorption on the support may be performed at low ionic strength, only lipases are adsorbed on these materials, enabling couple immobilization to the purification of the lipases [143]. Moreover, as the immobilization involves the open form of the lipase, a hyperactivated lipase form is obtained [143,213]. This adsorbed form of the lipases is more stable than the lipase in conformational equilibrium. It even is more stable than multipoint covalently immobilized lipases [214,215,216]. The properties of the immobilized enzymes on hydrophobic supports may be tuned by altering the support features (internal morphology, hydrophobicity, texture) [217,218] or the immobilization conditions [219,220,221], as it has been shown that the strong adsorption is enough to preserve the conformational form of the lipase induced by the immobilization conditions [220,221]. This way, a certain modulation of the properties of the final biocatalysts may be achieved using this reversible immobilization strategy. Using just one support and one enzyme, the final biocatalyst features (stability, activity, specificity) become very different depending on the immobilization conditions [219,220,221]. Thus, the control of the immobilization conditions must be quite strict to permit reproducible results. Lipase immobilization via interfacial activation on hydrophobic supports is a reversible process, enabling the reuse of the support after the enzyme immobilization [143]. That way, the global costs of the immobilization step may be reduced: the simplicity, increase in activity, and the possibility of support reuse are advantages from an economic point of view [154].
The main drawback of this immobilization strategy is founded on the reversibility of the immobilization: under certain circumstances, there are risks of enzyme release from the support [154,222]. This occurs using high temperature, high concentrations of hydrophobic cosolvents [223], or detergent-like reactants [222]. Researchers have proposed different solutions to this drawback. Physical cross-linking with ionic polymers maintains reversibility and can greatly reduce enzyme leakage [224,225]. Using covalent reactive polymers, such as polyethylenimine or aldehyde-dextran, avoids undesired enzyme release and ensures biocatalyst support reuse [226,227,228,229,230,231]. This makes the immobilization process more complex, demanding an additional reagent (the polymer) and an additional step in the biocatalyst preparation. Another alternative to prevent lipase release during operation is the use of heterofunctional supports, that is, supports bearing several functionalities able to interact with the enzyme [223,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247]. The combination of ionic and hydrophobic moieties on the support surface maintains the reversibility and reduces the risks of enzyme release [244]. This can make the support more expensive but maintains the simplicity of the immobilization protocol. The use of chemically reactive groups able to give enzyme-support covalent bonds can suppress enzyme release but also prevents support reuse [243,244,245]. Moreover, the immobilization protocol may be more complex, requiring more reagents and additional immobilization steps (e.g., some reagents and steps to suppress the chemical reactivity of the support after immobilization). However, this opens the opportunity for a final tailoring of the lipase features [235,236,238,243] and also for the co-immobilization of lipases bearing different stabilities with the possibility of reusing the most stable one [248,249,250].
An inadequate selection of the mechanical properties of the support used to immobilize the enzyme may have very negative effects on the operational stability of the biocatalysts. For example, the shear stress of stirrers may be able to break the support particles, and this has many negative effects on the operation performance [5,12,95]. The formation of fine solid particles leads to the blocking of the filters, making it necessary to discard the biocatalysts even when the enzyme is fully active. Another significant problem of using immobilized lipases is the possibility of the formation of a water layer inside the biocatalysts if the activity of the biocatalysts is high enough [59,250,251]. That way, even if some strategy is utilized to control the bulk water activity, and a system using very low water activity is employed, this water layer around the enzyme can lead to enzyme inactivation, for example, in continuous oil transesterification [252]. The alcohol and mainly the acid may become concentrated in this aqueous phase, reducing the pH value where the enzyme is located and exposing the enzyme to very high alcohol concentrations. Furthermore, short-chain acids and alcohols may inactivate the enzyme and are also strong inhibitors acting as competitive inhibitors [13,14,16,17,37,144,253,254]. This can be at least partially solved using very hydrophobic supports, where water will not become adsorbed, or using ultrasounds, able to stir inside the particles and preventing the water layer formation [179,180,252,255,256,257,258,259,260,261,262]. For example, the use of Novozym® 435, a commercial immobilized preparation of the lipase B from Candida antarctica using the interfacial adsorption of the enzyme in a moderately hydrophobic acrylic resin, Lewatit® VP OC 1600 (a macroporous support formed by poly(methyl methacrylate) cross-linked with divinylbenzene) and perhaps the most used immobilized lipase biocatalysts present all these limitations: the mechanical resistance is short, the support permits the formation of a hydrophobic layer compromising the enzyme stability, enzyme leakage can occur, the support can be even dissolved in some solvents due to a no proper cross-linking [5,263].
Therefore, not only a proper immobilization protocol must be adopted but many other aspects must be considered to prepare a suitable industrial lipase biocatalyst.

2.5. Current Status of Immobilized Lipases on Short-Chain Ester Syntheses

The literature describes many efforts to produce short-chain esters using immobilized lipases. Some of them are focused on the chemical reaction, its optimization, and the understanding of kinetics and thermodynamics aspects, but a considerable part of them is focused on the immobilization of lipases, whose efficacy is tested in short-chain ester syntheses. Generally, these last ones do not include assays on a wide range of reaction conditions. However, these studies are important in the general context of the enzymatic ester syntheses, due to the aforementioned drawbacks related to the biocatalytic route, and the relevance of immobilization of enzymes for industrial applications.
Table 1 and Table 2 present a bibliographic survey on the main reaction parameters for the enzymatic syntheses of esters via direct esterification with chain length of up to 10 carbons in the final molecule.

2.6. Performance Indicators on Enzymatic Esterification Studies and Environmental Aspects

Few studies achieved productivities higher than 100 g L−1 h−1 in enzymatic short-chain ester syntheses. The highest result was obtained by Garcia-Cebrian et al. (2018) [291], 472.1 and 259.1 g L−1 h−1, in solvent-free ethyl valerate synthesis using microwave and conventional heating, respectively. Other studies also found productivities close to 100 g L−1 h−1, such as Jaiswal and Rathod (2018) [277] in isobutyl propionate manufacture, Musa et al. (2018) [296] in ethyl hexanoate synthesis, and Friedrich et al. (2012) [182] in ethyl butyrate production. The main features of these studies are the short reaction time (1–3 h) and high product concentrations obtained. In this sense, we emphasize two important aspects—at least one of the substrates has a chain length equal to or superior to 4, which means that one of them is not so strong an inhibitor/inactivator of the acting lipases, and high biocatalyst loadings (superior to 5% w/w) are utilized. For comparative purposes, the work of Jiang and co-authors used a copper salt as a heterogeneous catalyst in the synthesis of isoamyl acetate resulting in a productivity of 358.9 g L−1 h−1 [86]. The biocatalyst loading is directly correlated with reaction time, but high loadings have operational and economical constraints for the process, as previously discussed. The long reaction time (higher than 24 h) required for some syntheses [278,279,289,293] may be associated with enzymatic inhibition or inactivation caused by short-chain acids and alcohols even using solvents. Thus, the use of solvents in this reaction is justified to reduce the concentration of substrates as they are potential inhibitors/inactivating agents for lipases. These reagents are relatively hydrophilic, which means that they can interact with the solvation layer of the enzyme altering the three-dimensional enzyme structure [14,15,16,302,303,304,305,306]. This may affect the required flexibility of the enzymes, and consequently, their catalytic activity [60,303,307,308,309]. Therefore, the inhibition/inactivating potential and the hydrophilic characteristic of short-chain acids and alcohols are significant challenges for the optimum performance of immobilized lipases. Some authors evaluate the enzymatic activity assays using standard substrates for hydrolysis, that are also different from the ones in the esterification reaction (for instance, p-nitrophenyl derivatives, olive oil, or oleic acid esters); this hydrolytic activity may not be directly related to the esterification activity [178,293,296,310]. There are some possible approaches to circumvent these drawbacks: (i) the adoption of high biocatalyst loadings (>5% w/w), aiming to increase reaction rates; (ii) optimization of the immobilization protocol and a careful choice of immobilization support to improve enzyme activity and stability.
It is possible to observe from the data in Table 1 and Table 2 that concentrations of obtained esters below 50 g L−1 are frequent, which is low for industrial applications. This result is directly associated with the use of the solvents. Even in fermentative processes, that are known for low product concentrations and the need for complex downstream operations to separate the main product and co-products of cell metabolism, concentrations of products superior to 50 g L−1 are usually obtained [11,311]. Therefore, it is hard to justify in economic terms the adoption of expensive catalysts such as immobilized lipases for producing commodity/specialty chemicals in the presence of solvents and at low concentrations. On the other hand, concentrations as high as 100 g L−1 using solvents were achieved by some authors [178,272,286,300]. An important aspect to highlight in these studies is the possibility of working with high concentrations of substrates. For instance, Anschau and co-workers (2021) [286] used Lipozyme® TL-IM as a biocatalyst in the synthesis of isopentyl butyrate in n-hexane without any water removal method and using response surface methodologies (RSM) as an optimization strategy; relatively high acid concentrations were used (0.5 mol L−1 of butyric acid), indicating the significant tolerance of this lipase for butyric acid [286]. However, this result was achieved with high biocatalyst loading (11.5 g L−1 as reported, equivalent to 21.3% w/w total substrate mass) and a relatively long reaction time (6 h). Thus, it is possible to obtain high concentrations of the product even using solvents, since high substrate concentrations are tolerated by the immobilized lipase. Fed-batch strategies are also interesting alternatives to circumvent the effects on immobilized lipases of high concentrations of substrates.
Considering KPI—yield, concentration, and productivity—it is possible to suggest how far short-chain ester syntheses using biocatalysis are from industrial applications. Using the feasibility criteria proposed by Tüvfesson and co-workers (2011) [8] in a biocatalytic production—2000 kg product per one kg of immobilized biocatalyst for specialties chemicals—we can check that even high concentrations of products result in low kg product per kg biocatalyst, as the results reported by Musa et al. (2018) [296]—only 4.2 kg of product per kg of biocatalyst. Some works [21,108,194,278,279,280,293] argue that biocatalysis may bring advantages in terms of simplification of processes (and consequential reduction in costs) due to the specificity and selectivity of biocatalysts, but the need for solvent separation and recovery in a biocatalytic process will represent the opposite.
The acid-mediated syntheses (using H2SO4 or other inorganic ones) are generally carried out for short reaction times, varying from 1 to 8 h in many cases [2]; contrarily, a significant part of the studies shown in Table 1 and Table 2 present reaction times equal to or higher than 24 h, although many of them did not include a detailed evaluation of the reaction time; this permits to speculate that equilibrium may be achieved earlier than reported. The long reaction times bring economic constraints related to the increase in operational and labor costs, as well as a reduction in the productivity of the process. The lack of explicit information about the initial quantities of reagents used and/or the volume of the reaction media impairs a reliable calculation of concentrations of products and biocatalyst productivity. An interesting study carried out by Kovalenko et al. (2021) [312] synthesized many short-chain esters using in-house immobilized lipases, but this study did not give specific details on the yields for each product. A proper investigation of economic constraints on this topic depends on these measures. However, it is important to highlight that many studies described in Table 1 and Table 2 delve into exploring some scientific aspects, independently from a potential scaling-up and the development of a biocatalytic process.
The importance of a proper exploration of the economic aspects of biocatalytic production relies on the environmental advantages that can be obtained from enzymatic technology for ester production. Thum and Oxenbøll (2008) [9] present a comparison between esters obtained in chemical and biocatalytic routes from an environmental perspective—energy savings, reduction in greenhouse gas emissions, reduction in the generated wastewater, non-hazardous production (in the case of solvent-free systems), among other important findings. These environmental advantages associated with higher purity of obtained products, due to the high specificity of immobilized lipases as catalysts and mild reaction conditions, may lead to products with differentiated prices that reach specific high added-value markets. Furthermore, these advantages may be the drivers pushing chemical industries to increase their sustainability indicators. Biocatalytic routes to produce important compounds such as short-chain esters may have a range of positive market appeals as clean and sustainable production (related to the Sustainable Development Goals from the United Nations), Green Chemistry, Bioeconomy, and energy transition. In summary, it is worth further exploring the enzymatic short-chain ester syntheses aiming to overcome the technical and economic drawbacks.

2.7. Solvents and Solvent-Free Systems on Enzymatic Esterification Studies

The use of hydrophobic solvents is disseminated in most of the studies, particularly hydrocarbons. Hydrocarbons are frequently miscible with low concentrations of acids and alcohols and are highly volatile, favoring the purification of the products. Another important feature of the solvents in esterification reactions is to promote an increase in the pKa of the acids, and only non-ionized forms are valid for the chemical equilibrium of esterification reactions [17,313,314]. On the other hand, hydrocarbon solvents are toxic and flammable bringing safety and health concerns, in addition to other disadvantages associated with the key performance indicators (KPI), such as the decreased concentration of the product and productivity. n-Hexane, cyclohexane, n-heptane, or toluene, are considered problematic or hazardous compounds by the CHEM21 selection guide of solvents [75]. When solvents are indispensable to perform a reaction, green solvents should be the selected alternative. Green solvents are characterized by their low environmental and safety impact during production, use, and final destination [315,316]. In this context, green solvents are compounds that conciliate solvent technical features (stability, inertness, viscosity, volatility) with a green/renewable obtaining process, non-toxicity, and biodegradability. Green solvents such as ionic liquids, deep-eutectic solvents, supercritical CO2, and other bio-based compounds have been intensively studied in the past years [159,314,315,316]. A survey on the Web of Science database revealed 1084 scientific papers with “green solvent” as a keyword in the past 5 years. Furthermore, some short-chain esters like ethyl acetate, isopropyl acetate, butyl acetate, and amyl acetate, are also considered green solvents in CHEM21 classification [75]. Other esters, such as lactones, seem promising solvent alternatives for biocatalytic oxidation reactions [317]. However, we have found few enzymatic short-chain esters studies that use green solvents [254,273].
Even using solvents, a significant part of the studies reached the highest carboxylic acid transformation to ester when using a stoichiometric excess of the alcohol higher than 3, as may be seen in Table 1 and Table 2. This excess of alcohol also leads to low concentrations of products. Moreover, in some cases, the high stoichiometric excesses of alcohol may suggest that there is a combination of the solvent and the surplus alcohol. For example, Baek and co-workers (2020) [264] found that phenomenon when 0.9 mols L−1 of n-octanol was used utilizing 1,2-dichloroethane as solvent. This solvent is only moderately hydrophobic (logP 1.48), but the overall hydrophobicity can be improved by the presence of the surplus n-octanol (logP 3.0).
In solvent-free systems (SFSs), there is a great potential increase in product concentration, as demonstrated by Musa et al. (2018) [296], even if lower conversions than in the presence of solvent are achieved (47.0% versus 69.0%) or by the study of Salah and co-workers (2007) [203], as can be seen in Table 1. Other solvent-free syntheses also obtained concentrations of products higher than 100 g L−1 [37,178,275,277,291,298,299]. The syntheses carried out in SFSs permit different strategies to improve the process. The most obvious is the adoption of stoichiometric excesses of alcohol, taking as an example the study of Lorenzo et al. (2022) [21] in the synthesis of isopentyl acetate, and Dai et al. (2014) [275] in the synthesis of butyl propionate. Isopentanol and n-butanol, in these cases, act simultaneously as reagents and solvents. However, in this kind of approach, two important consequences should be evaluated—the formed water (if not removed from the media) is miscible with the surplus alcohol, and the surplus alcohol will increase the probability of enzyme inhibition. The occurrence of inhibition by excess alcohol will cause an increase in the reaction time, demanding high biocatalyst loadings to compensate for the reduction in reaction rates. This effect has a drastic impact on the productivity of the syntheses, as can be observed in some studies [37,175,275].
Some studies also consider the addition of water in the reaction media, in different proportions, as well as keeping the produced water in the reaction media. At first glance, these approaches seem counterintuitive considering that water accumulation will favor hydrolysis. However, as highly hydrophilic reagents may cause distortions in the hydration layer of lipases, the addition of water may compensate for these deleterious effects [17,37]. Thus, the presence of water causes simultaneously opposite impacts—a positive one, in kinetic aspects, and a negative one, from a thermodynamic perspective [18]. Another possible negative impact, discussed by Krishna and Karanth (2001) [318] and other authors is the deactivation (partial or total) of the lipases due to the acidification of the microaqueous environment of the enzyme, considering that short-chain acids have relatively low pKa [17]. The impact of the addition of water on the yield of esters varies greatly, as can be seen in Table 1 [37,61,178,183,203,292]. There is a great dependence on the amount of water, the reaction media, and the immobilized lipase studied, considering that each immobilized lipase has an optimum water activity [111].

2.8. The Use of Ultrasound and Microwave on Enzymatic Esterification Studies

One way to improve solvent-free short-chain ester production is the use of microwaves or ultrasound [276]. As both technologies enhance the reaction rates, and the last one prevents the formation of a water layer inside the biocatalyst particles, the effects of reagents inhibition (or even partial inactivation) caused by the short-chain reagents on reaction rates become less pronounced [261,277]. These technologies, despite intense lab-scale exploration, are still far from application in the industrial context.
The use of ultrasound promotes a very effective agitation, even inside the particles (that way preventing the hydrophilic phase formation in the enzyme environment) and improve the mixing of phases, which is crucial to avoid any mass transfer limitations when immobilized enzymes are present [1,261]. Microwaves, on the other hand, act under kinetic reaction conditions due to the orientation of polar functional groups of the reagents and improve enzyme catalytic activity and stability [1,290]. Ultrasonically assisted synthesis, also known as sonochemistry, is an emerging technology that has been employed in the production of esters. This technique is based on the generation of high-frequency sound waves (>16 kHz) in a liquid medium. These sound waves induce vibrations that facilitate the movement of reagent molecules, thereby increasing interactions among them. Additionally, ultrasound can trigger cavitation, a phenomenon that creates localized supercritical conditions characterized by high temperatures and pressures [1,261]. Cavitation occurs when high-frequency sound waves generate vapor bubbles within the liquid. These bubbles grow and collapse rapidly, generating shear forces and impact that drive the progression of the enzymatic reaction and improve the solubility of the reagents. The intensity of ultrasound is a crucial factor that affects the efficacy of ester synthesis through this technique. Excessively high intensities can damage the enzyme structure, while excessively low intensities may not induce sufficient cavitation [277]. Several experimental studies have confirmed the effectiveness of ultrasound to improve ester synthesis, and here the work of Jaiswal and Rathod (2018) [277] is a successful example of its use, obtaining high ester concentration and productivity in the synthesis of isobutyl propionate in SFSs.
The other alternative to improve the reaction is the use of microwave irradiation. It covers a wide range of frequencies and wavelengths (1 mm to 1 m and 300 GHz to 300 MHz) [319]. This technology involves essential components such as the wave source, transmission lines, and applicator [297]. The effectiveness of microwave irradiation heating depends on the dielectric properties of the material, penetration depth, frequency, and mode of operation (monomode or multimode) [320,321,322]. Monomode reactors accurately direct radiation to the sample, maximizing energy efficiency, while multimode reactors distribute radiation chaotically, resulting in non-uniform heating. In addition to thermal effects, the scientific literature debates the possibility of non-thermal effects related to the direct interaction of the electric field with specific molecules [320]. Microwave technology offers several advantages, including ease of use, energy savings, reduced processing time, and a lower alcohol ratio to obtain similar yields in ester production. Some studies using conventional catalysts, such as that of Choedkiatsahul et al. (2015) [323], demonstrated a 99.4% conversion in just 1.75 min at 70 °C with the addition of 1% NaOH when using microwaves in the production of esters from palm oil. Similarly, Yu et al. (2010) [324] carried out a noteworthy study on the synthesis of methyl esters of fatty acids using microwaves, achieving a 94% conversion in 12 h, in contrast to the 24 h required in the conventional method. In the field of biocatalytic synthesis, where challenges like low reaction rates, long reaction periods, and low conversions are common, the application of microwaves stands out as an efficient solution. Microwave irradiation provides precise heating, substantially increasing the reaction rate and significantly reducing synthesis reaction time [325]. Furthermore, microwaves have a greater impact on molecules with higher dipole moments, such as lipases and substrates used in enzymatic synthesis with polar properties, making the process more effective by facilitating interaction between the substrate and the biocatalyst [325]. Some notable examples of enzymatic short-chain ester syntheses using microwaves are described in Table 1, in the works of Cebrian-Garcia (2018) [291] and Bhavsar and Yadav (2018) [276].

2.9. Biocatalysts on Enzymatic Esterifications and Reusability Studies

Different biocatalysts were used in the studies. Novozym® 435, a widely studied commercial lipase from Novozymes®, was used in many studies as can be seen in Table 1 and Table 2, even with the problems this biocatalyst presents. Its high catalytic activity and stability in different chemical environments and temperatures are crucial for synthetic applications; the main features, advantages, and disadvantages of Novozym® 435 were in-depth reviewed in the work of Ortiz and co-authors (2019) [5]. Novozym® 435 permits to reach high ester yields for the syntheses of octyl formate, hexyl formate, phenethyl formate [264,297], butyl propionate [275,276], butyl butyrate [278], and 3-hexen-1-yl acetate [274].
Novozym® 435 shows better performance than other biocatalysts in ester syntheses, including comparisons with other immobilized forms of lipase B of Candida antarctica [12,42,183,276]. This is also the case for short-chain ester syntheses, as shown in Table 1 [182,183,276]. One of the main reasons is its stability and high catalytic activity in more hydrophilic and acidic environments. Hydrophobic solvents are generally adopted for short-chain ester syntheses, as already discussed. Still, Novozym® 435 shows good performance in the presence of hydrophilic solvents, as demonstrated in the work of Hasegawa and co-authors (2008) in the synthesis of ethyl lactate in 1,4-dioxane [326]. Moreover, lactic acid is a relatively strong organic acid (pKa equal to 3.86, lower than acetic acid, 4.76). Novozym® 435 stability in such environments allows solvent-free short-chain ester syntheses [15,21,268,274,275,276,327].
The main limitation of Novozym® 435 is relative to its immobilization support. This support is poly(methyl methacrylate) cross-linked with divinylbenzene and commercially denominated Lewatit® VP OC 1600. Lewatit® VP OC 1600 has a surface area of 130 m2 g−1, an average particle size of 315–1000 μm and pore diameter of 150 Å, (product information Lewatit® VP OC 1600, Lanxess, edition: 13 October 2011). The support is moderately hydrophobic and permits enzyme immobilization via interfacial activation, with the advantages previously described. However, its only moderate hydrophobicity makes relatively simple the release of the enzyme during operation [328,329,330,331,332,333]. Furthermore, it facilitates the retention of water and hydrophilic substrates in the enzyme environment, which as stated above is a major reason for the enzyme inactivation in esterification processes [59,252,260,266,334,335]. The support presents relatively high mechanical fragility that causes problems for the industrial operation of the biocatalysts [5]. An even more serious problem is that the cross-linking seems to be not fully efficient, and the support can be dissolved in certain solvents, causing problems not only related to the biocatalyst operation stability, but also to the downstream [336,337]. Even with these problems, this biocatalyst remains the most used lipase biocatalyst in academic literature.
As a significant part of the studies are focused on the immobilization of lipases, and the short-chain ester synthesis is just one test to characterize the biocatalyst features, a variety of lipases and immobilization support materials are observed in Table 1 and Table 2. In short-chain ester syntheses there are hydrophilic environments and the formation of water; thus, highly hydrophobic support tends to have better performance. Highly hydrophilic supports tend to favor the water phase formation in the lipase environment, favoring the hydrolysis of the formed ester if water is not removed/absorbed, as in the case of ethyl hexanoate synthesis mediated by lipase immobilized onto chitosan [296]. Another possible effect of the support nature is the dissolution of highly hydrophilic support (and resultant enzyme release) in the reaction media if high temperatures are present [5,154]. A recent article from Bolivar and co-authors (2022) [154] argues that immobilization is not yet a mature discipline due to some misconceptions in the development of the support, the immobilization protocol adopted, and its application in a specific reaction media/reactor.
Another possible approach in enzymatic ester synthesis is the utilization of aggregated lipases (e.g., lyophilized or precipitated) as a method for enzyme immobilization. The use of powder enzymes is one of the oldest “immobilization” techniques, but it only works under conditions where enzyme solubilization is not possible, for example, in organic solvents, as the enzyme is not soluble in these media. As this immobilized lipase form is not supported in any material, this represents a lower cost in biocatalyst preparation [111,299]. Generally, these biocatalysts may present low enzymatic activity, but strategies such as salt-induced activation may improve their activity [111,338]. Meneses and co-authors found that lyophilized CALB had similar performance (>90% conversion) than Novozym® 435 in the synthesis of benzyl propionate in a fed-batch system [298]; methyl benzoate was synthesized by Leszczak and Tran-Minh using lyophilized Candida rugosa lipase suspended in hexane-toluene [339]; ethyl butyrate, ethyl hexanoate, and ethyl caprylate were also synthesized by lyophilized Rhizopus chinensis lipase at 40 °C achieving conversions close to or higher than 90% after 20 h in the work of Sun and co-workers (2009) [340].
The biocatalyst loadings used for short-chain ester syntheses vary greatly between the different studies, including the way it is represented in the different publications (concentration, mass, activity, etc.), making their comparison a complex task in many instances. In this review, we have expressed the biocatalyst loading as the total mass (enzyme + immobilization support) for practical reasons; really, the total biocatalyst load influences the reactor and processes design. Another alternative, perhaps reflecting better the enzyme performance, would have been the use of the used amount of enzymes, but that was impossible to calculate in most of the publications, mainly if they used commercial biocatalysts. It was possible to observe high biocatalyst loadings in many studies (superior to 5% w/w or equivalent measures) from Table 1 and Table 2. High loadings are required in most cases due to accentuated inhibition and partial deactivation effects of short-chain reagents on the performance of immobilized lipases, and this brings two obstacles for scaling up of a process: (i) additional costs, and (ii) operational issues due to mixing and effective mass transfer in continuous stirred tank reactors or fluidized reactors [12,341]. Immobilized lipases are still a high-cost income and thus, high loadings mean a drastic increase in the operational costs of the process [8,12,66]. Furthermore, there is no proportional correlation between the increase in biocatalyst loading and reusability [12]. In terms of agitation, high loadings of catalysts tend to decrease the operational stability of immobilization supports, due to the potential increase in high-energy collisions between biocatalyst particles [5,12,37,342]. This leads to a decrease in the reusability of immobilized lipases with consequential impacts on operational costs. Moreover, if a high percentage of the reactor volume is the biocatalyst particles, a large amount of the final reaction mixture will be retained inside the particles and between particles, requiring extensive washing to recover it.
One of the advantages of the biocatalytic route using immobilized lipases is the possibility of reuse of the biocatalyst, due to the easy removal of a solid catalyst from a liquid reaction media. The maximum number of sequential cycles of reaction for a given biocatalyst has the utmost importance for the determination of the production cost per batch. Reusability is affected by the nature of the immobilization support, its resistance (chemical and mechanical), and the severity of the reaction media, which includes temperature, reaction time, and agitation intensity [5,12,159]. Few studies dedicated an in-depth evaluation of this crucial aspect. Generally, the topic is limited to checking how many reaction cycles (in optimized conditions) it is possible to carry out without significant loss of enzymatic activity. The number of reuses is frequently inferior to 10 cycles using different immobilized lipases in most studies, as shown in Table 3.
There are a few exceptions to this general lack of long-time operational biocatalyst stability, such as the studies of Dai et al. [25], Seo et al. [267], Pires-Cabral et al. [278], and Martins et al. [259]. This means that, in general, we lack information on the immobilization support materials‘ behavior when submitted to standard mechanical/orbital agitation in the presence of organic solvents for long reaction times.

3. Conclusions

Short-chain esters have a huge relevance in the chemical industry context and technologies for their current production are well established. However, the market sizes and the different sectors involved (particularly the ones that deal with daily products) may be a driver for the gradual implementation of new technologies that reduce environmental impacts and greenhouse gas emissions. The scientific community has been proposing alternatives to short-chain esters production in a greener and more sustainable way. The adoption of immobilized lipases as biocatalysts in these processes may be a very suitable solution from these points of view, but it has technical and economic obstacles to be overcome, particularly the necessity of using organic solvents and the high inhibition potential of the substrates on the lipases, which lead to unfavorable process performance indicators. Further developments in enzymatic immobilization, the understanding of thermodynamics and kinetics aspects of the esterification processes, and the establishment of effective reaction setups are crucial to obtaining cost-effective and efficient biocatalysts and the full conversion of the substrates into products at high rates. Preliminary techno-economic analyses are scarce in the literature for biocatalytic ester production, and this information is of utmost importance to establish reliable quantitative metrics on the lab-scale to be used in the future scaling-up of the processes. This gap of information should be filled to take advantage of the high positive environmental indicators that biocatalytic routes present when compared to conventional chemical routes.

Author Contributions

R.R.d.S., M.M.d.S. and M.W.R.M.: Writing—original draft, Conceptualization. Á.B.-M., E.A.M., D.M.G.F., V.S.F.-L. and R.F.-L.: Writing—review and editing. D.M.G.F., V.S.F.-L. and R.F.-L.: Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We gratefully recognize the financial support from Fundação Carlos Chagas Filho de Amparo à Pesquisa do Estado do Rio de Janeiro—FAPERJ (APQ1 E26/211.736/2021 and IT Grant E-26/260.418/2022), Ministerio de Ciencia e Innovación and Agencia Estatal de Investigación (Spanish Government) (PID2022-136535OB-I00). A.B.M. thanks the Generalitat Valenciana (CIPROM/2021/70) for financial support and PID2021-123079OB-I00 project funded by MCIN/AEI/10.13039/501100011033 and “ERDF A way of making Europe”.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Khan, N.R.; Rathod, V.K. Enzyme catalyzed synthesis of cosmetic esters and its intensification: A review. Process Biochem. 2015, 50, 1793–1806. [Google Scholar] [CrossRef]
  2. Khan, Z.; Javed, F.; Shamair, Z.; Hafeez, A.; Fazal, T.; Aslam, A.; Zimmerman, W.B.; Rehman, F. Current developments in esterification reaction: A review on process and parameters. J. Ind. Eng. Chem. 2021, 103, 80–101. [Google Scholar] [CrossRef]
  3. Sá, A.G.A.; de Meneses, A.C.; de Araújo, P.H.H.; de Oliveira, D. A review on enzymatic synthesis of aromatic esters used as flavor ingredients for food, cosmetics and pharmaceuticals industries. Trends Food Sci. Technol. 2017, 69, 95–105. [Google Scholar] [CrossRef]
  4. Nielsen, P.M.; Brask, J.; Fjerbaek, L. Enzymatic biodiesel production: Technical and economical considerations. Eur. J. Lipid Sci. Technol. 2008, 110, 692–700. [Google Scholar] [CrossRef]
  5. Ortiz, C.; Ferreira, M.L.; Barbosa, O.; Dos Santos, J.C.S.; Rodrigues, R.C.; Berenguer-Murcia, Á.; Briand, L.E.; Fernandez-Lafuente, R. Novozym 435: The “perfect” lipase immobilized biocatalyst? Catal. Sci. Technol. 2019, 9, 2380–2420. [Google Scholar] [CrossRef]
  6. Manjón, A.; Iborra, J.L.; Arocas, A. Short-chain flavour ester synthesis by immobilized lipase in organic media. Biotechnol. Lett. 1991, 13, 339–344. [Google Scholar] [CrossRef]
  7. Wu, S.; Snajdrova, R.; Moore, J.C.; Baldenius, K.; Bornscheuer, U.T. Biocatalysis: Enzymatic Synthesis for Industrial Applications. Angew. Chem. Int. Ed. 2021, 60, 88–119. [Google Scholar] [CrossRef]
  8. Tufvesson, P.; Lima-Ramos, J.; Nordblad, M.; Woodley, J.M. Guidelines and cost analysis for catalyst production in biocatalytic processes. Org. Process Res. Dev. 2011, 15, 266–274. [Google Scholar] [CrossRef]
  9. Thum, O.; Oxenbøll, K.M. Biocatalysis—A Sustainable Method for the Production of Emollient Esters. SÖFW-Journal 2008, 134, 44–47. [Google Scholar]
  10. Basso, A.; Serban, S. Industrial applications of immobilized enzymes—A review. Mol. Catal. 2019, 479, 110607. [Google Scholar] [CrossRef]
  11. Straathof, A.J.J. Transformation of biomass into commodity chemicals using enzymes or cells. Chem. Rev. 2014, 114, 1871–1908. [Google Scholar] [CrossRef] [PubMed]
  12. De Sousa, R.R.; Costa, M.; Pinto, C.; Cristina, E.; Aguieiras, G.; Cipolatti, E.P.; Andrade, E.; Ayla, M.; Ana, S.; Pinto, J.C.; et al. Comparative performance and reusability studies of lipases on syntheses of octyl esters with an economic approach. Bioprocess Biosyst. Eng. 2022, 45, 131–145. [Google Scholar] [CrossRef] [PubMed]
  13. Bélafi-Bakó, K.; Badr, A.K.; Ehrenstein, U.; Gubicza, L. Kinetics of Ethyl Acetate Formation by Lipase in Organic Solvent and Solvent-Free System. Chem. Pap. 2003, 57, 278–281. [Google Scholar]
  14. Bezbradica, D.; Mijin, D.; Šiler-Marinković, S.; Knežević, Z. The effect of substrate polarity on the lipase-catalyzed synthesis of aroma esters in solvent-free systems. J. Mol. Catal. B Enzym. 2007, 45, 97–101. [Google Scholar] [CrossRef]
  15. Nordblad, M.; Adlercreutz, P. Effects of acid concentration and solvent choice on enzymatic acrylation by Candida antarctica lipase B. J. Biotechnol. 2008, 133, 127–133. [Google Scholar] [CrossRef]
  16. Hari Krishna, S.; Karanth, N.G. Lipases and Lipase-Catalyzed Esterification Reactions in Nonaqueous Media. Catal. Rev. 2002, 44, 499–591. [Google Scholar] [CrossRef]
  17. Sousa, R.R.; Silva, A.S.; Fernandez-Lafuente, R.; Ferreira-Leitão, V.S. Solvent-free esterifications mediated by immobilized lipases: A review from thermodynamic and kinetics perspective. Catal. Sci. Technol. 2021, 11, 5696–5711. [Google Scholar] [CrossRef]
  18. de Sousa, R.R.; da Silva, A.S.; Fernandez-Lafuente, R.; Ferreira-Leitão, V.S. Simplified method to optimize enzymatic esters syntheses in solvent-free systems: Validation using literature and experimental data. Catalysts 2021, 11, 1357. [Google Scholar] [CrossRef]
  19. Gao, P.; Jiang, Q.; Xu, Y.; Xia, W. Biosynthesis of acetate esters by dominate strains, isolated from Chinese traditional fermented fish (Suan yu). Food Chem. 2018, 244, 44–49. [Google Scholar] [CrossRef]
  20. Gao, W.; Wu, K.; Chen, L.; Fan, H.; Zhao, Z.; Gao, B.; Wang, H.; Wei, D. A novel esterase from a marine mud metagenomic library for biocatalytic synthesis of short-chain flavor esters. Microb. Cell Fact. 2016, 15, 41. [Google Scholar] [CrossRef]
  21. Lorenzo, B.; Fernández, L.; Ortega, J.; Domínguez, L. Improvements in the Modeling and Kinetics Processes of the Enzymatic Synthesis of Pentyl Acetate. Processes 2023, 11, 1640. [Google Scholar] [CrossRef]
  22. Brault, G.; Shareck, F.; Hurtubise, Y.; Lépine, F.; Doucet, N. Short-Chain Flavor Ester Synthesis in Organic Media by an E. coli Whole-Cell Biocatalyst Expressing a Newly Characterized Heterologous Lipase. PLoS ONE 2014, 9, e91872. [Google Scholar] [CrossRef]
  23. Kruis, A.J.; Bohnenkamp, A.C.; Patinios, C.; Van Nuland, Y.M.; Levisson, M.; Mars, A.E.; Van Den Berg, C.; Kengen, S.W.M.; Weusthuis, R.A. Microbial production of short and medium chain esters: Enzymes, pathways, and applications. Biotechnol. Adv. 2019, 37, 107407. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, C.; Liang, X.; Abdo, A.A.A.; Kaddour, B.; Li, X.; Teng, C.; Wan, C. Ultrasound-assisted lipase-catalyzed synthesis of ethyl acetate: Process optimization and kinetic study. Biotechnol. Biotechnol. Equip. 2021, 35, 255–263. [Google Scholar] [CrossRef]
  25. Mehta, A.; Grover, C.; Bhardwaj, K.K.; Gupta, R. Application of lipase purified from Aspergillus fumigatus in the syntheses of ethyl acetate and ethyl lactate. J. Oleo Sci. 2020, 69, 23–29. [Google Scholar] [CrossRef]
  26. Ferone, M.; Raganati, F.; Olivieri, G. Bioreactors for succinic acid production processes. Crit. Rev. Biotechnol. 2019, 39, 571–586. [Google Scholar] [CrossRef]
  27. Barel, A.O.; Paye, M.; Maibach, H.I. (Eds.) Handbook of Cosmetic Science and Technology, 3rd ed.; Informa Healthcare: London, UK, 2010; Volume 36, ISBN 9781420069631. [Google Scholar]
  28. Vogel, A.I. 130. Physical properties and chemical constitution. Part XIII. Aliphatic carboxylic esters. J. Chem. Soc. 1948, 624–644. [Google Scholar] [CrossRef]
  29. Kononenko, O.K.; Herstein, K.M. Nonaqueous Solvents for Sucrose. Ind. Eng. Chem. Chem. Eng. Data Ser. 1956, 1, 87–92. [Google Scholar] [CrossRef]
  30. Prat, D.; Hayler, J.; Wells, A. A survey of solvent selection guides. Green Chem. 2014, 16, 4546–4551. [Google Scholar] [CrossRef]
  31. Welton, T. Solvents and sustainable chemistry. Proc. R. Soc. A Math. Phys. Eng. Sci. 2015, 471, 20150502. [Google Scholar] [CrossRef]
  32. International Organization of the Flavor Industry. IOFI Code of Practices; IOFI: Washington, DC, USA, 2012; Volume 1.3. [Google Scholar]
  33. Belitz, H.-D.; Grosch, W.; Schieberle, P. Food Chemistry; Springer: Berlin/Heidelberg, Germany, 2008; ISBN 9783540699330. [Google Scholar]
  34. Sun, J.; Liu, S.Q. Ester Synthesis in Aqueous Media by Lipase: Alcoholysis, Esterification and Substrate Hydrophobicity. J. Food Biochem. 2015, 39, 11–18. [Google Scholar] [CrossRef]
  35. Sun, J.; Yu, B.; Curran, P.; Liu, S.Q. Lipase-catalysed transesterification of coconut oil with fusel alcohols in a solvent-free system. Food Chem. 2012, 134, 89–94. [Google Scholar] [CrossRef]
  36. Kuperkar, V.V.; Lade, V.G.; Prakash, A.; Rathod, V.K. Synthesis of isobutyl propionate using immobilized lipase in a solvent free system: Optimization and kinetic studies. J. Mol. Catal. B Enzym. 2014, 99, 143–149. [Google Scholar] [CrossRef]
  37. Chowdary, G.V.; Ramesh, M.N.; Prapulla, S.G. Enzymic synthesis of isoamyl isovalerate using immobilized lipase from Rhizomucor miehei: A multivariate analysis. Process Biochem. 2000, 36, 331–339. [Google Scholar] [CrossRef]
  38. Ergan, F.; Trani, M.; André, G. Use of lipases in multiphasic systems solely composed of substrates. J. Am. Oil Chem. Soc. 1991, 68, 412–417. [Google Scholar] [CrossRef]
  39. Fellechner, O.; Blatkiewicz, M.; Smirnova, I. Reactive Separations for In Situ Product Removal of Enzymatic Reactions: A Review. Chemie-Ingenieur-Technik 2019, 91, 1522–1543. [Google Scholar] [CrossRef]
  40. Lee, D.E.; Park, K.M.; Choi, S.J.; Shim, J.H.; Chang, P.S. Enhancing operational stability and exhibition of enzyme activity by removing water in the immobilized lipase-catalyzed production of erythorbyl laurate. Biotechnol. Prog. 2013, 29, 882–889. [Google Scholar] [CrossRef]
  41. Vadgama, R.N.; Odaneth, A.A.; Lali, A.M. Green synthesis of isopropyl myristate in novel single phase medium Part I: Batch optimization studies. Biotechnol. Rep. 2015, 8, 133–137. [Google Scholar] [CrossRef]
  42. Sandoval, G.; Condoret, J.S.; Monsan, P.; Marty, A. Esterification by Immobilized Lipase in Solvent-Free Media: Kinetic and Thermodynamic Arguments. Biotechnol. Bioeng. 2002, 78, 313–320. [Google Scholar] [CrossRef]
  43. Castillo, E.; Dossat, V.; Marty, A.; Stéphane Condoret, J.; Combes, D. The role of silica gel in lipase-catalyzed esterification reactions of high-polar substrates. JAOCS J. Am. Oil Chem. Soc. 1997, 74, 77–85. [Google Scholar] [CrossRef]
  44. Kobayashi, T.; Adachi, S. Reaction equilibrium for lipase-catalyzed condensation in organic solvent systems. Biotechnol. Lett. 2004, 26, 1461–1468. [Google Scholar] [CrossRef] [PubMed]
  45. Sousa, R.R.; Pazutti, L.V.B.; Dalmaso, G.Z.L.; Siqueira, D.F.; Silva, A.S.; Ferreira-Leitão, V.S. A practical approach to obtain high yield lipase-mediated synthesis of octyl caprylate with Novozym 435. Biocatal. Biotransform. 2020, 38, 293–303. [Google Scholar] [CrossRef]
  46. Sose, M.T.; Bansode, S.R.; Rathod, V.K. Solvent free lipase catalyzed synthesis of butyl caprylate. J. Chem. Sci. 2017, 129, 1755–1760. [Google Scholar] [CrossRef]
  47. Santos, J.C.; Bueno, T.; Rós, P.C.M.; de Castro, H.F. Lipase-catalyzed Synthesis of Butyl Esters by Direct Esterification in Solvent-Free System. J. Chem. Technol. Biotechnol. 2007, 82, 956–961. [Google Scholar] [CrossRef]
  48. Gubicza, L.; Kabiri-Badr, A.; Keoves, E.; Belafi-Bako, K. Large-scale enzymatic production of natural flavour esters in organic solvent with continuous water removal. J. Biotechnol. 2000, 84, 193–196. [Google Scholar] [CrossRef]
  49. Woodcock, L.L.; Wiles, C.; Greenway, G.M.; Watts, P.; Wells, A.; Eyley, S. Enzymatic synthesis of a series of alkyl esters using Novozym 435 in a packed-bed, miniaturized, continuous flow reactor. Biocatal. Biotransform. 2008, 26, 466–472. [Google Scholar] [CrossRef]
  50. Reis, C.; Yvay, E.; De Sousa, A.; De França, J.; Casemiro, R. Design of immobilzed enzyme biocatalyst: Drawbacks and opportinities. Quim. Nov. 2019, 42, 768–783. [Google Scholar]
  51. Serrano-Arnaldos, M.; Bastida, J.; Máximo, F.; Ortega-Requena, S.; Montiel, C. One-Step Solvent-Free Production of a Spermaceti Analogue Using Commercial Immobilized Lipases. ChemistrySelect 2018, 3, 748–752. [Google Scholar] [CrossRef]
  52. Gubicza, L.; Bélafi-Bakó, K.; Fehér, E.; Fráter, T. Waste-free process for continuous flow enzymatic esterification using a double pervaporation system. Green Chem. 2008, 10, 1284–1287. [Google Scholar] [CrossRef]
  53. Kanwar, S.S.; Verma, H.K.; Pathak, S.; Kaushal, R.K.; Kumar, Y.; Verma, M.L.; Chimni, S.S.; Chauhan, G.S. Enhancement of ethyl propionate synthesis by poly (AAc-co-HPMA-cl-MBAm)-immobilized Pseudomonas aeruginosa MTCC-4713, exposed to Hg2+ and NH4+ ions. Acta Microbiol. Immunol. Hung. 2006, 53, 195–207. [Google Scholar] [CrossRef]
  54. Foukis, A.; Gkini, O.A.; Stergiou, P.Y.; Papamichael, E.M. New insights and tools for the elucidation of lipase catalyzed esterification reaction mechanism in n-hexane: The synthesis of ethyl butyrate. Mol. Catal. 2018, 455, 159–163. [Google Scholar] [CrossRef]
  55. Santos, M.; Velez, A.; Abburra, R.; Magario, I. Thermodynamic and kinetic aspects of the solvent-free liquid–liquid synthesis of tailor-made glyceride esters. Fuel 2023, 333, 126461. [Google Scholar] [CrossRef]
  56. Hartman, L. Miscibility of glycerol with fatty acids and glycerides. J. Am. Oil Chem. Soc. 1963, 40, 142. [Google Scholar] [CrossRef]
  57. Dhake, K.P.; Thakare, D.D.; Bhanage, B.M. Lipase: A potential biocatalyst for the synthesis of valuable flavour and fragrance ester compounds. Flavour Fragr. J. 2013, 28, 71–83. [Google Scholar] [CrossRef]
  58. Foresti, M.L.; Pedernera, M.; Bucalá, V.; Ferreira, M.L. Multiple effects of water on solvent-free enzymatic esterifications. Enzym. Microb. Technol. 2007, 41, 62–70. [Google Scholar] [CrossRef]
  59. Marty, A.; Dossat, V.; Condoret, J.S. Continuous operation of lipase-catalyzed reactions in nonaqueous solvents: Influence of the production of hydrophilic compounds. Biotechnol. Bioeng. 1997, 56, 232–237. [Google Scholar] [CrossRef]
  60. Halling, P.J.; Finney, J.L.; Ho, M.W.; Franks, F.; Littlechild, J.A. What can we learn by studying enzymes in non-aqueous media? Philos. Trans. R. Soc. B Biol. Sci. 2004, 359, 1287–1297. [Google Scholar] [CrossRef]
  61. Ghamgui, H.; Karra-chaabouni, M.; Bezzine, S.; Miled, N.; Gargouri, Y. Production of isoamyl acetate with immobilized Staphylococcus simulans lipase in a solvent-free system. Enzym. Microb. Technol. 2006, 38, 788–794. [Google Scholar] [CrossRef]
  62. Foresti, M.L.; Pedernera, M.; Ferreira, M.L.; Bucalá, V. Kinetic modeling of enzymatic ethyl oleate synthesis carried out in biphasic systems. Appl. Catal. A Gen. 2008, 334, 65–72. [Google Scholar] [CrossRef]
  63. Foresti, M.L.; Ferreira, M.L. Solvent-free ethyl oleate synthesis mediated by lipase from Candida antarctica B adsorbed on polypropylene powder. Catal. Today 2005, 107–108, 23–30. [Google Scholar] [CrossRef]
  64. Vaysse, L.; Ly, A.; Moulin, G.; Dubreucq, E. Chain-length selectivity of various lipases during hydrolysis, esterification and alcoholysis in biphasic aqueous medium. Enzym. Microb. Technol. 2002, 31, 648–655. [Google Scholar] [CrossRef]
  65. Goldberg, M.; Thomas, D.; Legoy, M.D. Water activity as a key parameter of synthesis reactions: The example of lipase in biphasic (liquid/solid) media. Enzym. Microb. Technol. 1990, 12, 976–981. [Google Scholar] [CrossRef]
  66. de Sousa, R.R.; de Castro, R.d.P.V.; Assis, N.M.; da Silva, A.S.; Freire, D.M.G.; Fernandez-Lafuente, R.; Ferreira-Leitão, V.S. Technical–Economic Assessment—The Missing Piece for Increasing the Attractiveness of Applied Biocatalysis in Ester Syntheses? Catalysts 2023, 13, 223. [Google Scholar] [CrossRef]
  67. Clyburn, F. IFF Proxy Statement and Notice of 2022 Annual Meeting of Sharingholders. Available online: https://ir.iff.com/static-files/d874db63-45ff-4a28-8bf5-88b2d50ff719 (accessed on 12 August 2024).
  68. Ansorge-Schumacher, M.B.; Thum, O. Immobilised lipases in the cosmetics industry. Chem. Soc. Rev. 2013, 42, 6475–6490. [Google Scholar] [CrossRef] [PubMed]
  69. Europe, C.D. The In-Cosmetics Global Award Winners 2024. Available online: https://www.cosmeticsdesign-europe.com/Article/2024/04/22/The-In-Cosmetics-Global-award-winners-2024 (accessed on 31 July 2024).
  70. Cosmetic Business. The Winners of the Pure Beauty Global Awards 2024 Revealed. Available online: https://cosmeticsbusiness.com/pure-beauty-global-awards-winners-2024 (accessed on 3 January 2025).
  71. Barcelona Perfumery Congress. The Barcelona Perfumery Awards Already Have Their Winners Among 120 International Nominations from Across the Value Chain of the Sector. Available online: https://perfumerycongress.com/meet-the-winners-of-the-2023-barcelona-perfumery-awards/ (accessed on 3 January 2025).
  72. Future Market Insights Inc. Esters Market Outlook from 2023 to 2033. Available online: https://www.futuremarketinsights.com/reports/esters-market (accessed on 12 August 2024).
  73. Grand View Research U.S. Esters Market Size, Share & Trends Analysis Report By Product (Fatty Esters, Phosphate Esters, Acrylic Esters, Cellulose Esters, Allyl and Aromatic Esters), By Application, and Segment Forecasts, 2019–2025. Available online: https://www.grandviewresearch.com/industry-analysis/us-esters-market (accessed on 13 August 2024).
  74. Csanádi, Z.; Kurdi, R.; Bélafi-Bakó, K. Ethyl-Acetate Synthesis in Gas Phase by Immobilised Lipase. Hung. J. Ind. Chem. Veszpr. 2012, 40, 39–44. [Google Scholar]
  75. Prat, D.; Wells, A.; Hayler, J.; Sneddon, H.; McElroy, C.R.; Abou-Shehada, S.; Dunn, P.J. CHEM21 selection guide of classical- and less classical-solvents. Green Chem. 2016, 18, 288–296. [Google Scholar] [CrossRef]
  76. Mary Bailey. BASF to Expand Production Capacity of Synthetic Ester Base Stocks in China—Chemical Engineering. Available online: https://www.chemengonline.com/basf-to-expand-production-capacity-of-synthetic-ester-base-stocks-in-china/?printmode=1 (accessed on 4 May 2024).
  77. Wray, J. Evonik Grows Green Emollient Esters Business in China. Available online: https://cosmeticsbusiness.com/news/article_page/Evonik_grows_green_emollient_esters_business_in_China/201432 (accessed on 6 October 2023).
  78. Business Wire. Investment in Research and Development: OQ Chemicals Launches New Pilot Plant for Customized Esters. Business Wire. Available online: https://www.businesswire.com/news/home/20230504005125/en/Investment-in-Research-and-Development-OQ-Chemicals-Launches-New-Pilot-Plant-for-Customized-Esters (accessed on 5 February 2024).
  79. IFF. IFF Commissions Fragrance Ingredients Plant in China; Builds Industry Leadership; Sets New Sustainability Standards. International Flavors & Fragrances Inc. Available online: https://ir.iff.com/news-releases/news-release-details/iff-commissions-fragrance-ingredients-plant-china-builds-0 (accessed on 12 August 2024).
  80. Solvay. Solvay Opens Its Second Sustainable Solvent Plant in Brazil. Solvay. Available online: https://www.solvay.com/en/press-release/new-augeo-plant (accessed on 4 May 2024).
  81. Villa, C.; Mariani, E.; Loupy, A.; Grippo, C.; Grossi, G.C.; Bargagna, A. Solvent-free reactions as green chemistry procedures for the synthesis of cosmetic fatty esters. Green Chem. 2003, 5, 623–626. [Google Scholar] [CrossRef]
  82. Ishida, T.; Ogihara, Y.; Ohashi, H.; Akita, T.; Honma, T.; Oji, H.; Haruta, M. Base-free direct oxidation of 1-octanol to octanoic acid and its octyl ester over supported gold catalysts. ChemSusChem 2012, 5, 2243–2248. [Google Scholar] [CrossRef]
  83. Maruyama, S.A.; Kanda, L.R.S.; Wypych, F. Isopropyl octanoate synthesis catalyzed by layered zinc n-octanoate. J. Braz. Chem. Soc. 2017, 28, 985–994. [Google Scholar] [CrossRef]
  84. Jyoti, G.; Keshav, A.; Anandkumar, J.; Bhoi, S. Homogeneous and Heterogeneous Catalyzed Esterification of Acrylic Acid with Ethanol: Reaction Kinetics and Modeling. Int. J. Chem. Kinet. 2018, 50, 370–380. [Google Scholar] [CrossRef]
  85. Jiang, H. Reaction-controlled recovery of the copper(II) methanesulfonate catalyst for esterification. React. Kinet. Catal. Lett. 2005, 84, 223–228. [Google Scholar] [CrossRef]
  86. Zang, Y.; Shi, J.; Zhang, F.; Zhong, Y.; Zhu, W. Sulfonic acid-functionalized MIL-101 as a highly recyclable catalyst for esterification. Catal. Sci. Technol. 2013, 3, 2044–2049. [Google Scholar] [CrossRef]
  87. Park, Y.-M.; Lee, J.Y.; Chung, S.-H.; Park, I.S.; Lee, S.-Y.; Kim, D.-K.; Lee, J.-S.; Lee, K.-Y. Esterification of used vegetable oils using the heterogeneous WO3/ZrO2 catalyst for production of biodiesel. Bioresour. Technol. 2010, 101, S59–S61. [Google Scholar] [CrossRef]
  88. Ben Mya, O.; Bita, M.; Louafi, I.; Djouadi, A. Esterification process catalyzed by ZSM-5 zeolite synthesized via modified hydrothermal method. MethodsX 2018, 5, 277–282. [Google Scholar] [CrossRef]
  89. Yang, Z.; Zhao, L.; Lei, Z. Quaternary Ammonium Salt Functionalized Methoxypolyethylene Glycols-Supported Phosphotungstic Acid Catalyst for the Esterification of Carboxylic Acids with Alcohols. Catal. Lett. 2014, 144, 585–589. [Google Scholar] [CrossRef]
  90. Homaei, A.A.; Sariri, R.; Vianello, F.; Stevanato, R. Enzyme immobilization: An update. J. Chem. Biol. 2013, 6, 185–205. [Google Scholar] [CrossRef] [PubMed]
  91. Mohammadi Ziarani, G.; Lashgari, N.; Badiei, A. Sulfonic acid-functionalized mesoporous silica (SBA-Pr-SO3H) as solid acid catalyst in organic reactions. J. Mol. Catal. A Chem. 2015, 397, 166–191. [Google Scholar] [CrossRef]
  92. Poppe, J.K.; Fernandez-Lafuente, R.; Rodrigues, R.C.; Ayub, M.A.Z. Enzymatic reactors for biodiesel synthesis: Present status and future prospects. Biotechnol. Adv. 2015, 33, 511–525. [Google Scholar] [CrossRef] [PubMed]
  93. Contente, M.L.; Tamborini, L.; Molinari, F.; Paradisi, F. Aromas flow: Eco-friendly, continuous, and scalable preparation of flavour esters. J. Flow Chem. 2020, 10, 235–240. [Google Scholar] [CrossRef]
  94. Thum, O. Enzymatic production of care specialities based on fatty acid esters. Tenside Surfactants Deterg. 2004, 41, 287–290. [Google Scholar] [CrossRef]
  95. Baum, S.; Mueller, J.J.; Hilterhaus, L.; Eckstein, M.; Thum, O.; Liese, A. The bubble column reactor: A novel reactor type for cosmetic esters. In Applied Biocatalysis: From Fundamental Science to Industrial Applications; Wiley Online Library: Hoboken, NJ, USA, 2016; pp. 343–366. [Google Scholar] [CrossRef]
  96. Juan, J.C.; Zhang, J.; Yarmo, M.A. Efficient Esterification of Fatty Acids with Alcohols Catalyzed by Zr(SO4)2·4H2O under Solvent-Free Condition. Catal. Lett. 2008, 126, 319–324. [Google Scholar] [CrossRef]
  97. Devaraj Naik, B.; Udayakumar, M. Optimization studies on esterification of waste cooking oil using sulfated montmorillonite clay acidic catalyst. Mater. Today Proc. 2021, 46, 9855–9861. [Google Scholar] [CrossRef]
  98. Cao, M.; Peng, L.; Xie, Q.; Xing, K.; Lu, M.; Ji, J. Sulfonated Sargassum horneri carbon as solid acid catalyst to produce biodiesel via esterification. Bioresour. Technol. 2021, 324, 124614. [Google Scholar] [CrossRef]
  99. Mustafa, A.; Niikura, F. Green synthesis of isopropyl palmitate using immobilized Candida antarctica lipase: Process optimization using response surface methodology. Clean. Eng. Technol. 2022, 8, 100516. [Google Scholar] [CrossRef]
  100. Ögmundarson, Ó.; Sukumara, S.; Herrgård, M.J.; Fantke, P. Combining Environmental and Economic Performance for Bioprocess Optimization. Trends Biotechnol. 2020, 38, 1203–1214. [Google Scholar] [CrossRef]
  101. Delgove, M.A.F.; Laurent, A.B.; Woodley, J.M.; De Wildeman, S.M.A.; Bernaerts, K.V.; van der Meer, Y. A Prospective Life Cycle Assessment (LCA) of Monomer Synthesis: Comparison of Biocatalytic and Oxidative Chemistry. ChemSusChem 2019, 12, 1349–1360. [Google Scholar] [CrossRef] [PubMed]
  102. Serrano-Arnaldos, M.; Ortega-Requena, S.; Montiel, M.C.; Máximo, F.; Bastida, J.; Murcia, M.D. Preliminary economic assessment: A valuable tool to establish biocatalytic process feasibility with an in-lab immobilized lipase. J. Chem. Technol. Biotechnol. 2019, 94, 409–417. [Google Scholar] [CrossRef]
  103. Burton, S.G.; Cowan, D.A.; Woodley, J.M. The search for the ideal biocatalyst. Nat. Biotechnol. 2002, 20, 37–45. [Google Scholar] [CrossRef]
  104. Coelho, A.L.S.; Orlandelli, R.C. Immobilized microbial lipases in the food industry: A systematic literature review. Crit. Rev. Food Sci. Nutr. 2021, 61, 1689–1703. [Google Scholar] [CrossRef]
  105. Serrano-Arnaldos, M.; Montiel, M.C.; Ortega-Requena, S.; Máximo, F.; Bastida, J. Development and economic evaluation of an eco-friendly biocatalytic synthesis of emollient esters. Bioprocess Biosyst. Eng. 2020, 43, 495–505. [Google Scholar] [CrossRef]
  106. Martins, A.B.; Da Silva, A.M.; Schein, M.F.; Garcia-Galan, C.; Záchia Ayub, M.A.; Fernandez-Lafuente, R.; Rodrigues, R.C. Comparison of the performance of commercial immobilized lipases in the synthesis of different flavor esters. J. Mol. Catal. B Enzym. 2014, 105, 18–25. [Google Scholar] [CrossRef]
  107. Hills, G. Industrial use of lipases to produce fatty acid esters. Eur. J. Lipid Sci. Technol. 2003, 105, 601–607. [Google Scholar] [CrossRef]
  108. López-Fernández, J.; Benaiges, M.D.; Sebastian, X.; Bueno, J.M.; Valero, F. Producing Natural Flavours from Isoamyl Alcohol and Fusel Oil by Using Immobilised Rhizopus oryzae Lipase. Catalysts 2022, 12, 639. [Google Scholar] [CrossRef]
  109. Raghavendra, T.; Panchal, N.; Divecha, J.; Shah, A.; Madamwar, D. Biocatalytic synthesis of flavor ester “pentyl valerate” using Candida rugosa lipase immobilized in microemulsion based organogels: Effect of parameters and reusability. BioMed Res. Int. 2014, 2014, 353845. [Google Scholar] [CrossRef] [PubMed]
  110. Sheldon, R.A.; Woodley, J.M. Role of Biocatalysis in Sustainable Chemistry. Chem. Rev. 2018, 118, 801–838. [Google Scholar] [CrossRef]
  111. Adlercreutz, P. Immobilisation and application of lipases in organic media. Chem. Soc. Rev. 2013, 42, 6406–6436. [Google Scholar] [CrossRef]
  112. Gunukula, S.; Runge, T.; Anex, R. Assessment of Biocatalytic Production Parameters to Determine Economic and Environmental Viability. ACS Sustain. Chem. Eng. 2017, 5, 8119–8126. [Google Scholar] [CrossRef]
  113. Enzymes REACH Consortium. Welcome to Enzymes REACH Consortium. Available online: http://www.enzymes-reach.org/content/welcome-enzymes-reach-consortium (accessed on 6 January 2025).
  114. AMFEP. Major Regulatory Aspects on Technical Enzymes. Available online: https://amfep.org/about-enzymes/regulatory/technical-enzymes-regulations/ (accessed on 6 January 2025).
  115. Enzymes REACH Consortium. Safety Evaluation of Technical Enzyme Products with Regards to the REACH Legislation. 2021. Available online: http://www.enzymes-reach.org/content/documents (accessed on 6 January 2025).
  116. Melo, A.D.Q.; Silva, F.F.M.; Dos Santos, J.C.S.; Fernández-Lafuente, R.; Lemos, T.L.G.; Dias Filho, F.A. Synthesis of benzyl acetate catalyzed by lipase immobilized in nontoxic chitosan-polyphosphate beads. Molecules 2017, 22, 2165. [Google Scholar] [CrossRef]
  117. Jawale, P.V.; Bhanage, B.M. Synthesis of propyl benzoate by solvent-free immobilized lipase-catalyzed transesterification: Optimization and kinetic modeling. Bioprocess Biosyst. Eng. 2021, 44, 369–378. [Google Scholar] [CrossRef]
  118. Yadav, G.D.; Devendran, S. Lipase catalyzed synthesis of cinnamyl acetate via transesterification in non-aqueous medium. Process Biochem. 2012, 47, 496–502. [Google Scholar] [CrossRef]
  119. Sun, J.; Yu, B.; Curran, P.; Liu, S.Q. Lipase-catalysed ester synthesis in solvent-free oil system: Is it esterification or transesterification? Food Chem. 2013, 141, 2828–2832. [Google Scholar] [CrossRef]
  120. Verma, M.L.; Azmi, W.; Kanwar, S.S. Synthesis of ethyl acetate employing celite-immobilized lipase of Bacillus cereus MTCC 8372. Acta Microbiol. Immunol. Hung. 2009, 56, 229–242. [Google Scholar] [CrossRef] [PubMed]
  121. Kim, H.; Park, C. Enzymatic synthesis of phenethyl ester from phenethyl alcohol with acyl donors. Enzym. Microb. Technol. 2017, 100, 37–44. [Google Scholar] [CrossRef]
  122. Gómez, J.L.; Gómez, M.; Murcia, M.D.; Gómez, E.; Hidalgo, A.M.; Montiel, C.; Martínez, R. Biosynthesis of benzyl acetate: Optimization of experimental conditions, kinetic modelling and application of alternative methods for parameters determination. Bioresour. Technol. Rep. 2020, 11, 100519. [Google Scholar] [CrossRef]
  123. Chen, M.L.; Vali, S.R.; Lin, J.Y.; Ju, Y.H. Synthesis of the structured lipid 1,3-dioleoyl-2-palmitoylglycerol from palm oil. JAOCS J. Am. Oil Chem. Soc. 2004, 81, 525–532. [Google Scholar] [CrossRef]
  124. Palla, C.A.; Carrín, M.E. Kinetics modeling of the acidolysis with immobilized Rhizomucor miehei lipases for production of structured lipids from sunflower oil. Biochem. Eng. J. 2014, 90, 184–194. [Google Scholar] [CrossRef]
  125. Hita, E.; Robles, A.; Camacho, B.; González, P.A.; Esteban, L.; Jiménez, M.J.; Muñío, M.M.; Molina, E. Production of structured triacylglycerols by acidolysis catalyzed by lipases immobilized in a packed bed reactor. Biochem. Eng. J. 2009, 46, 257–264. [Google Scholar] [CrossRef]
  126. Jiménez, M.J.; Esteban, L.; Robles, A.; Hita, E.; González, P.A.; Muñío, M.M.; Molina, E. Production of triacylglycerols rich in palmitic acid at sn-2 position by lipase-catalyzed acidolysis. Biochem. Eng. J. 2010, 51, 172–179. [Google Scholar] [CrossRef]
  127. Gandhi, N.N.; Mukherjee, K.D. Reactivity of medium-chain substrates in the interesterification of tripalmitin catalyzed by papaya lipase. JAOCS J. Am. Oil Chem. Soc. 2001, 78, 965–968. [Google Scholar] [CrossRef]
  128. Subhedar, P.B.; Gogate, P.R. Ultrasound assisted intensification of biodiesel production using enzymatic interesterification. Ultrason. Sonochem. 2016, 29, 67–75. [Google Scholar] [CrossRef]
  129. Xie, W.; Zang, X. Immobilized lipase on core-shell structured Fe3O4-MCM-41 nanocomposites as a magnetically recyclable biocatalyst for interesterification of soybean oil and lard. Food Chem. 2016, 194, 1283–1292. [Google Scholar] [CrossRef] [PubMed]
  130. Zhang, S.-Y.; Fu, G.-N.; Du, L.-H.; Lin, H.; Zhang, A.-Y.; Xie, H.-J.; Sheng, Z.-K.; Xue, M.-M.; Yan, B.-L.; Liu, Y.; et al. Continuous flow biocatalysis: Synthesis of purine nucleoside esters catalyzed by lipase TL IM from Thermomyces lanuginosus. RSC Adv. 2024, 14, 10953–10961. [Google Scholar] [CrossRef] [PubMed]
  131. Wang, X.; Wang, X.; Wang, T. An effective method for reducing free fatty acid content of high-acid rice bran oil by enzymatic amidation. J. Ind. Eng. Chem. 2017, 48, 119–124. [Google Scholar] [CrossRef]
  132. Zhang, L.; Li, F.; Wang, C.; Zheng, L.; Wang, Z.; Zhao, R.; Wang, L. Lipase-mediated amidation of anilines with 1,3-diketones via C–C bond cleavage. Catalysts 2017, 7, 115. [Google Scholar] [CrossRef]
  133. Chen, Y.; Wang, Y.; Jin, Q.; Wang, X.; Hu, J.; Wang, X. Preparation of arachidonoyl ethanolamide by enzymatic amidation of arachidonic acid purified from microbial oil. Process Biochem. 2018, 66, 120–125. [Google Scholar] [CrossRef]
  134. Du, L.; Zheng, L.; Pan, Y.; Sheng, Z.; Zhang, S.; Lin, H.; Zhang, A.; Xie, H.; Luo, X. Highly Efficient Synthesis of Cinnamamides from Methyl Cinnamates and Phenylethylamines Catalyzed by Lipozyme® TL IM under Continuous-Flow Microreactors. Catalysts 2022, 12, 1265. [Google Scholar] [CrossRef]
  135. Zhang, A.-Y.; Huang, Z.-H.; Du, L.-H.; Lin, H.; Xie, H.-J.; Yan, B.-L.; Xue, M.M.; Wang, L.; Shao, W.-X.; Fu, G.-N.; et al. Pyrazine derivative synthesis in a continuous-flow system: A green synthesis of pyrazinamide from pyrazine esters and amines catalyzed by Lipozyme® TL IM from Thermomyces lanuginosus. RSC Adv. 2024, 14, 39560–39568. [Google Scholar] [CrossRef]
  136. Kapoor, M.; Gupta, M.N. Lipase promiscuity and its biochemical applications. Process Biochem. 2012, 47, 555–569. [Google Scholar] [CrossRef]
  137. Dwivedee, B.P.; Soni, S.; Sharma, M.; Bhaumik, J.; Laha, J.K.; Banerjee, U.C. Promiscuity of Lipase-Catalyzed Reactions for Organic Synthesis: A Recent Update. ChemistrySelect 2018, 3, 2441–2466. [Google Scholar] [CrossRef]
  138. Verger, R. “Interfacial activation” of lipases: Facts and artifacts. Trends Biotechnol. 1997, 15, 32–38. [Google Scholar] [CrossRef]
  139. Mancheño, J.M.; Pernas, M.A.; Martínez, M.J.; Ochoa, B.; Rúa, M.L.; Hermoso, J.A. Structural insights into the lipase/esterase behavior in the Candida rugosa lipases family: Crystal structure of the lipase 2 isoenzyme at 1.97A resolution. J. Mol. Biol. 2003, 332, 1059–1069. [Google Scholar] [CrossRef]
  140. Uppenberg, J.; Hansen, M.T.; Patkar, S.; Jones, T.A. The sequence, crystal structure determination and refinement of two crystal forms of lipase B from Candida antarctica. Structure 1994, 2, 293–308. [Google Scholar] [CrossRef]
  141. Rodrigues, R.C.; Fernandez-Lafuente, R. Lipase from Rhizomucor miehei as a biocatalyst in fats and oils modification. J. Mol. Catal. B Enzym. 2010, 66, 15–32. [Google Scholar] [CrossRef]
  142. Rodrigues, R.C.; Ortiz, C.; Berenguer-Murcia, Á.; Torres, R.; Fernández-Lafuente, R. Modifying enzyme activity and selectivity by immobilization. Chem. Soc. Rev. 2013, 42, 6290–6307. [Google Scholar] [CrossRef] [PubMed]
  143. Rodrigues, R.C.; Virgen-Ortíz, J.J.; dos Santos, J.C.S.; Berenguer-Murcia, Á.; Alcantara, A.R.; Barbosa, O.; Ortiz, C.; Fernandez-Lafuente, R. Immobilization of lipases on hydrophobic supports: Immobilization mechanism, advantages, problems, and solutions. Biotechnol. Adv. 2019, 37, 746–770. [Google Scholar] [CrossRef] [PubMed]
  144. Stergiou, P.Y.; Foukis, A.; Filippou, M.; Koukouritaki, M.; Parapouli, M.; Theodorou, L.G.; Hatziloukas, E.; Afendra, A.; Pandey, A.; Papamichael, E.M. Advances in lipase-catalyzed esterification reactions. Biotechnol. Adv. 2013, 31, 1846–1859. [Google Scholar] [CrossRef]
  145. Yu, H.; Qin, L.; Zhou, J. Effect of Oil Polarity on the Protein Adsorption at Oil-Water Interfaces. Langmuir 2023, 39, 10701–10710. [Google Scholar] [CrossRef]
  146. Stahmann, K.P.; Böddecker, T.; Sahm, H. Regulation and properties of a fungal lipase showing interfacial inactivation by gas bubbles, or droplets of lipid or fatty acid. Eur. J. Biochem. 1997, 244, 220–225. [Google Scholar] [CrossRef]
  147. Zhang, K.; Jin, Z.; Wang, P.; Zheng, S.-P.; Han, S.-Y.; Lin, Y. Improving the catalytic characteristics of lipase-displaying yeast cells by hydrophobic modification. Bioprocess Biosyst. Eng. 2017, 40, 1689–1699. [Google Scholar] [CrossRef]
  148. Wang, P.; He, J.; Sun, Y.; Reynolds, M.; Zhang, L.; Han, S.; Liang, S.; Sui, H.; Lin, Y. Display of fungal hydrophobin on the Pichia pastoris cell surface and its influence on Candida antarctica lipase B. Appl. Microbiol. Biotechnol. 2016, 100, 5883–5895. [Google Scholar] [CrossRef]
  149. Palomo, J.M.; Peñas, M.M.; Fernández-Lorente, G.; Mateo, C.; Pisabarro, A.G.; Fernández-Lafuente, R.; Ramírez, L.; Guisán, J.M. Solid-phase handling of hydrophobins: Immobilized hydrophobins as a new tool to study lipases. Biomacromolecules 2003, 4, 204–210. [Google Scholar] [CrossRef]
  150. Fernández-Lorente, G.; Palomo, J.M.; Fuentes, M.; Mateo, C.; Guisán, J.M.; Fernández-Lafuente, R. Self-assembly of Pseudomonas fluorescens lipase into bimolecular aggregates dramatically affects functional properties. Biotechnol. Bioeng. 2003, 82, 232–237. [Google Scholar] [CrossRef] [PubMed]
  151. Palomo, J.M.; Ortiz, C.; Fuentes, M.; Fernandez-Lorente, G.; Guisan, J.M.; Fernandez-Lafuente, R. Use of immobilized lipases for lipase purification via specific lipase-lipase interactions. J. Chromatogr. A 2004, 1038, 267–273. [Google Scholar] [CrossRef]
  152. Palomo, J.M.; Fuentes, M.; Fernández-Lorente, G.; Mateo, C.; Guisan, J.M.; Fernández-Lafuente, R. General trend of lipase to self-assemble giving bimolecular aggregates greatly modifies the enzyme functionality. Biomacromolecules 2003, 4, 1–6. [Google Scholar] [CrossRef] [PubMed]
  153. Wilson, L.; Palomo, J.M.; Fernández-Lorente, G.; Illanes, A.; Guisán, J.M.; Fernández-Lafuente, R. Effect of lipase–lipase interactions in the activity, stability and specificity of a lipase from Alcaligenes sp. Enzym. Microb. Technol. 2006, 39, 259–264. [Google Scholar] [CrossRef]
  154. Bolivar, J.M.; Woodley, J.M.; Fernandez-Lafuente, R. Is enzyme immobilization a mature discipline? Some critical considerations to capitalize on the benefits of immobilization. Chem. Soc. Rev. 2022, 51, 6251–6290. [Google Scholar] [CrossRef] [PubMed]
  155. Sigma-Aldrich. Lipase Liquid. Available online: https://www.sigmaaldrich.com/BR/pt/search/lipase-liquid?focus=products&page=1&perpage=30&sort=relevance&term=lipaseliquid&type=product (accessed on 6 January 2025).
  156. Collaço, A.C.A.; Aguieiras, E.C.G.; Santos, J.G.; de Oliveira, R.A.; de Paula Vieira de Castro, R.; Freire, D.M.G. Experimental study and preliminary economic evaluation of enzymatic biodiesel production by an integrated process using co-products from palm (Elaeais guineensis Jaquim) industry. Ind. Crops Prod. 2020, 157, 112904. [Google Scholar] [CrossRef]
  157. Girelli, A.M.; Astolfi, M.L.; Scuto, F.R. Agro-industrial wastes as potential carriers for enzyme immobilization: A review. Chemosphere 2020, 244, 125368. [Google Scholar] [CrossRef]
  158. Collaço, A.C.A.; Aguieiras, E.C.G.; Cavalcanti, E.D.C.; Freire, D.M.G. Development of an integrated process involving palm industry co-products for monoglyceride / diglyceride emulsifier synthesis: Use of palm cake and fiber for lipase production and palm fatty-acid distillate as raw material. LWT 2021, 135, 110039. [Google Scholar] [CrossRef]
  159. Aguieiras, E.C.G.; Ribeiro, D.S.; Couteiro, P.P.; Bastos, C.M.B.; de Queiroz, D.S.; Parreira, J.M.; Langone, M.A.P. Investigation of the Reuse of Immobilized Lipases in Biodiesel Synthesis: Influence of Different Solvents in Lipase Activity. Appl. Biochem. Biotechnol. 2016, 179, 485–496. [Google Scholar] [CrossRef]
  160. Kumar, A.; Dhar, K.; Kanwar, S.S.; Arora, P.K. Lipase catalysis in organic solvents: Advantages and applications. Biol. Proced. Online 2016, 18, 2. [Google Scholar] [CrossRef] [PubMed]
  161. Delavault, A.; Grüninger, J.; Kapp, D.; Hollenbach, R.; Rudat, J.; Ochsenreither, K.; Syldatk, C. Enzymatic Synthesis of Alkyl Glucosides by β-Glucosidases in a 2-in-1 Deep Eutectic Solvent System. Chemie-Ingenieur-Technik 2022, 94, 417–426. [Google Scholar] [CrossRef]
  162. Guajardo, N.; de María, P.D. Lipases in Green Chemistry: Deep Eutectic Solvents (DES) as New Green Solvents. In Lipases and Phospholipases: Methods and Protocols, Methods in Molecular Biology; Springer: Berlin/Heidelberg, Germany, 2018; Volume 1835, pp. 288–297. ISBN 9781493986729. [Google Scholar]
  163. Di Cosimo, R.; Mc Auliffe, J.; Poulose, A.J.; Bohlmann, G. Industrial use of immobilized enzymes. Chem. Soc. Rev. 2013, 42, 6437–6474. [Google Scholar] [CrossRef] [PubMed]
  164. Liese, A.; Hilterhaus, L. Evaluation of immobilized enzymes for industrial applications. Chem. Soc. Rev. 2013, 42, 6236–6249. [Google Scholar] [CrossRef]
  165. Sheldon, R.A.; van Pelt, S. Enzyme immobilisation in biocatalysis: Why, what and how. Chem. Soc. Rev. 2013, 42, 6223–6235. [Google Scholar] [CrossRef]
  166. Monteiro, R.R.C.; Arana-Peña, S.; da Rocha, T.N.; Miranda, L.P.; Berenguer-Murcia, Á.; Tardioli, P.W.; dos Santos, J.C.S.; Fernandez-Lafuente, R. Liquid lipase preparations designed for industrial production of biodiesel. Is it really an optimal solution? Renew. Energy 2021, 164, 1566–1587. [Google Scholar] [CrossRef]
  167. Iyer, P.V.; Ananthanarayan, L. Enzyme stability and stabilization—Aqueous and non-aqueous environment. Process Biochem. 2008, 43, 1019–1032. [Google Scholar] [CrossRef]
  168. Fernandez-Lafuente, R. Stabilization of multimeric enzymes: Strategies to prevent subunit dissociation. Enzym. Microb. Technol. 2009, 45, 405–418. [Google Scholar] [CrossRef]
  169. Rodrigues, R.C.; Berenguer-Murcia, Á.; Carballares, D.; Morellon-Sterling, R.; Fernandez-Lafuente, R. Stabilization of enzymes via immobilization: Multipoint covalent attachment and other stabilization strategies. Biotechnol. Adv. 2021, 52, 107821. [Google Scholar] [CrossRef]
  170. Garcia-Galan, C.; Berenguer-Murcia, Á.; Fernandez-Lafuente, R.; Rodrigues, R.C. Potential of different enzyme immobilization strategies to improve enzyme performance. Adv. Synth. Catal. 2011, 353, 2885–2904. [Google Scholar] [CrossRef]
  171. Mateo, C.; Palomo, J.M.; Fernandez-Lorente, G.; Guisan, J.M.; Fernandez-Lafuente, R. Improvement of enzyme activity, stability and selectivity via immobilization techniques. Enzym. Microb. Technol. 2007, 40, 1451–1463. [Google Scholar] [CrossRef]
  172. Guimarães, J.R.; Miranda, L.P.; Fernandez-Lafuente, R.; Tardioli, P.W. Immobilization of Eversa® Transform via CLEA Technology Converts It in a Suitable Biocatalyst for Biolubricant Production Using Waste Cooking Oil. Molecules 2021, 26, 193. [Google Scholar] [CrossRef] [PubMed]
  173. Kumari, A.; Mahapatra, P.; Kumar, G.V.; Banerjee, R. Comparative study of thermostabilty and ester synthesis ability of free and immobilized lipases on cross linked silica gel. Bioprocess Biosyst. Eng. 2008, 31, 291–298. [Google Scholar] [CrossRef]
  174. Junior, J.M.; Mattos, F.R.; Costa, G.R.; Zurlo, A.B.R.; Fernandez-Lafuente, R.; Mendes, A.A. Improved Catalytic Performance of Lipase Eversa® Transform 2.0 via Immobilization for the Sustainable Production of Flavor Esters—Adsorption Process and Environmental Assessment Studies. Catalysts 2022, 12, 1412. [Google Scholar] [CrossRef]
  175. Kaur, P.; Jana, A.K.; Jana, M.M. Immobilization of Candida rugosa lipase on optimized polyamidoamine dendrimer functionalized magnetic multiwalled carbon nanotubes for green manufacture of butyl butyrate ester. Mol. Catal. 2024, 553, 113779. [Google Scholar] [CrossRef]
  176. Badgujar, K.C.; Bhanage, B.M. The combine use of ultrasound and lipase immobilized on co-polymer matrix for efficient biocatalytic application studies. J. Mol. Catal. B Enzym. 2015, 122, 255–264. [Google Scholar] [CrossRef]
  177. Da Silva, V.C.F.; Contesini, F.J.; Carvalho, P.D.O. Characterization and catalytic activity of free and immobilized lipase from Aspergillus niger: A comparative study. J. Braz. Chem. Soc. 2008, 19, 1468–1474. [Google Scholar] [CrossRef]
  178. Barbosa, O.; Ortiz, C.; Berenguer-Murcia, Á.; Torres, R.; Rodrigues, R.C.; Fernandez-Lafuente, R. Strategies for the one-step immobilization-purification of enzymes as industrial biocatalysts. Biotechnol. Adv. 2015, 33, 435–456. [Google Scholar] [CrossRef]
  179. Yang, J.; Ma, X.; Zhang, Z.; Chen, B.; Li, S.; Wang, G. Lipase immobilized by modification-coupled and adsorption-cross-linking methods: A comparative study. Biotechnol. Adv. 2010, 28, 644–650. [Google Scholar] [CrossRef]
  180. Pavlidis, I.V.; Tsoufis, T.; Enotiadis, A.; Gournis, D.; Stamatis, H. Functionalized multi-wall carbon nanotubes for lipase immobilization. Adv. Eng. Mater. 2010, 12, 179–183. [Google Scholar] [CrossRef]
  181. Graebin, N.G.; Martins, A.B.; Lorenzoni, A.S.G.; Garcia-Galan, C.; Fernandez-Lafuente, R.; Ayub, M.A.Z.; Rodrigues, R.C. Immobilization of lipase B from Candida antarctica on porous styrene-divinylbenzene beads improves butyl acetate synthesis. Biotechnol. Prog. 2012, 28, 406–412. [Google Scholar] [CrossRef]
  182. Friedrich, J.L.R.; Peña, F.P.; Garcia-Galan, C.; Fernandez-Lafuente, R.; Ayub, M.A.Z.; Rodrigues, R.C. Effect of immobilization protocol on optimal conditions of ethyl butyrate synthesis catalyzed by lipase B from Candida antarctica. J. Chem. Technol. Biotechnol. 2012, 88, 1089–1095. [Google Scholar] [CrossRef]
  183. Basri, M.; Wong, C.C.; Ahmad, M.B.; Razak, C.N.A.; Salleh, A.B. Immobilization of lipase on poly(N-vinyl-2-pyrrolidone-co-2-hydroxyethyl methacrylate) hydrogel for the synthesis of butyl oleate. JAOCS J. Am. Oil Chem. Soc. 1999, 76, 571–577. [Google Scholar] [CrossRef]
  184. Vaidya, L.B.; Nadar, S.S.; Rathod, V.K. Entrapment of surfactant modified lipase within zeolitic imidazolate framework (ZIF)-8. Int. J. Biol. Macromol. 2020, 146, 678–686. [Google Scholar] [CrossRef] [PubMed]
  185. Batista, K.A.; Lopes, F.M.; Yamashita, F.; Fernandes, K.F. Lipase entrapment in PVA/Chitosan biodegradable film for reactor coatings. Mater. Sci. Eng. C Mater. Biol. Appl. 2013, 33, 1696–1701. [Google Scholar] [CrossRef]
  186. Kheadr, E.E.; Vuillemard, J.-C.; El-Deeb, S.A. Acceleration of Cheddar Cheese Lipolysis by Using Liposome-entrapped Lipases. J. Food Sci. 2002, 67, 485–492. [Google Scholar] [CrossRef]
  187. Reetz, M.T.; Zonta, A.; Simpelkamp, J. Efficient immobilization of lipases by entrapment in hydrophobic sol-gel materials. Biotechnol. Bioeng. 1996, 49, 527–534. [Google Scholar] [CrossRef]
  188. Sittko, I.; Kremser, K.; Roth, M.; Kuehne, S.; Stuhr, S.; Tiller, J.C. Amphiphilic polymer conetworks with defined nanostructure and tailored swelling behavior for exploring the activation of an entrapped lipase in organic solvents. Polymer 2015, 64, 122–129. [Google Scholar] [CrossRef]
  189. Li, Y.; Hu, M.; Du, Y.; Xiao, H.; McClements, D.J. Control of lipase digestibility of emulsified lipids by encapsulation within calcium alginate beads. Food Hydrocoll. 2011, 25, 122–130. [Google Scholar] [CrossRef]
  190. Nadar, S.S.; Rathod, V.K. Encapsulation of lipase within metal-organic framework (MOF) with enhanced activity intensified under ultrasound. Enzym. Microb. Technol. 2018, 108, 11–20. [Google Scholar] [CrossRef]
  191. Reetz, M.T.; Tielmann, P.; Wiesenhöfer, W.; Könen, W.; Zonta, A. Second Generation Sol-Gel Encapsulated Lipases: Robust Heterogeneous Biocatalysts. Adv. Synth. Catal. 2003, 345, 717–728. [Google Scholar] [CrossRef]
  192. Ozturk, T.K.; Kilinc, A. Immobilization of lipase in organic solvent in the presence of fatty acid additives. J. Mol. Catal. B Enzym. 2010, 67, 214–218. [Google Scholar] [CrossRef]
  193. Bayramoǧlu, G.; Hazer, B.; Altintaş, B.; Arica, M.Y. Covalent immobilization of lipase onto amine functionalized polypropylene membrane and its application in green apple flavor (ethyl valerate) synthesis. Process Biochem. 2011, 46, 372–378. [Google Scholar] [CrossRef]
  194. Mendes, A.A.; de Castro, H.F.; Giordano, R.L.C. Covalent attachment of lipases on glyoxyl-agarose beads: Application in fruit flavor and biodiesel synthesis. Int. J. Biol. Macromol. 2014, 70, 78–85. [Google Scholar] [CrossRef]
  195. Kartal, F. Enhanced esterification activity through interfacial activation and cross-linked immobilization mechanism of Rhizopus oryzae lipase in a nonaqueous medium. Biotechnol. Prog. 2016, 32, 899–904. [Google Scholar] [CrossRef]
  196. Sampaio, C.S.; Angelotti, J.A.F.; Fernandez-Lafuente, R.; Hirata, D.B. Lipase immobilization via cross-linked enzyme aggregates: Problems and prospects—A review. Int. J. Biol. Macromol. 2022, 215, 434–449. [Google Scholar] [CrossRef]
  197. Kanwar, S.S. Synthesis of Ethyl Propionate Catalyzed by Poly(NAEAAm-co-AAc)-cl-MBAm Hydrogel-Immobilized Lipase of Bacillus coagulans MTCC-6375. J. Appl. Polym. Sci. 2007, 116, 2658–2667. [Google Scholar] [CrossRef]
  198. Cipolatti, E.P.; Pinto, M.C.C.; Robert, J.d.M.; da Silva, T.P.; Beralto, T.d.C.; Santos, J.G.F.; de Castro, R.d.P.V.; Fernandez-Lafuente, R.; Manoel, E.A.; Pinto, J.C.; et al. Pilot-scale development of core–shell polymer supports for the immobilization of recombinant lipase B from Candida antarctica and their application in the production of ethyl esters from residual fatty acids. J. Appl. Polym. Sci. 2018, 135, 46727. [Google Scholar] [CrossRef]
  199. Silva, M.V.C.; Rosa, C.M.R.; Aguiar, L.G.; Oliveira, P.C.; de Castro, H.F.; Freitas, L. Synthesis of Isopropyl Palmitate by Lipase Immobilized on a Magnetized Polymer Matrix. Chem. Eng. Technol. 2020, 43, 1741–1748. [Google Scholar] [CrossRef]
  200. Cipolatti, E.P.; Valério, A.; Henriques, R.O.; Cerqueira Pinto, M.C.; Lorente, G.F.; Manoel, E.A.; Guisán, J.M.; Ninow, J.L.; de Oliveira, D.; Pessela, B.C. Production of new nanobiocatalysts via immobilization of lipase B from C. antarctica on polyurethane nanosupports for application on food and pharmaceutical industries. Int. J. Biol. Macromol. 2020, 165, 2957–2963. [Google Scholar] [CrossRef]
  201. Gawlitza, K.; Wu, C.; Georgieva, R.; Wang, D. Immobilization of lipase B within micron-sized poly- N -isopropylacrylamide hydrogel particles by solvent exchange. Phys. Chem. Chem. Phys. 2012, 14, 9594–9600. [Google Scholar] [CrossRef]
  202. Poojari, Y.; Beemat, J.S.; Clarson, S.J. Enzymatic synthesis of poly(ε-caprolactone): Thermal properties, recovery, and reuse of lipase B from Candida antarctica immobilized on macroporous acrylic resin particles. Polym. Bull. 2013, 70, 1543–1552. [Google Scholar] [CrossRef]
  203. Ben Salah, R.; Ghamghui, H.; Miled, N.; Mejdoub, H.; Gargouri, Y. Production of butyl acetate ester by lipase from novel strain of Rhizopus oryzae. J. Biosci. Bioeng. 2007, 103, 368–372. [Google Scholar] [CrossRef]
  204. Barbosa, O.; Ortiz, C.; Torres, R.; Fernandez-Lafuente, R. Effect of the immobilization protocol on the properties of lipase B from Candida antarctica in organic media: Enantiospecifc production of atenolol acetate. J. Mol. Catal. B Enzym. 2011, 71, 124–132. [Google Scholar] [CrossRef]
  205. Dosanjh, N.S.; Kaur, J. Immobilization, stability and esterification studies of a lipase from a Bacillus sp. Biotechnol. Appl. Biochem. 2002, 36, 7. [Google Scholar] [CrossRef] [PubMed]
  206. Ghamgui, H.; Karra-Chaabouni, M.; Gargouri, Y. 1-Butyl oleate synthesis by immobilized lipase from Rhizopus oryzae: A comparative study between n-hexane and solvent-free system. Enzym. Microb. Technol. 2004, 35, 355–363. [Google Scholar] [CrossRef]
  207. Shu, C.; Cai, J.; Huang, L.; Zhu, X.; Xu, Z. Biocatalytic production of ethyl butyrate from butyric acid with immobilized Candida rugosa lipase on cotton cloth. J. Mol. Catal. B Enzym. 2011, 72, 139–144. [Google Scholar] [CrossRef]
  208. Brígida, A.I.S.; Pinheiro, Á.D.T.; Ferreira, A.L.O.; Gonçalves, L.R.B. Immobilization of Candida antarctica lipase B by adsorption to green coconut fiber. Appl. Biochem. Biotechnol. 2008, 146, 173–187. [Google Scholar] [CrossRef]
  209. Ittrat, P.; Chacho, T.; Pholprayoon, J.; Suttiwarayanon, N.; Charoenpanich, J. Application of agriculture waste as a support for lipase immobilization. Biocatal. Agric. Biotechnol. 2014, 3, 77–82. [Google Scholar] [CrossRef]
  210. Bonet-Ragel, K.; López-Pou, L.; Tutusaus, G.; Benaiges, M.D.; Valero, F. Rice husk ash as a potential carrier for the immobilization of lipases applied in the enzymatic production of biodiesel. Biocatal. Biotransform. 2018, 36, 151–158. [Google Scholar] [CrossRef]
  211. de Castro, H.F.; de Lima, R.; Roberto, I.C. Rice Straw as a Support for Immobilization of Microbial Lipase. Biotechnol. Prog. 2001, 17, 1061–1064. [Google Scholar] [CrossRef] [PubMed]
  212. Abdulla, R.; Sanny, S.A.; Derman, E. Stability studies of immobilized lipase on rice husk and eggshell membrane. IOP Conf. Ser. Mater. Sci. Eng. 2017, 206, 012032. [Google Scholar] [CrossRef]
  213. Manoel, E.A.; dos Santos, J.C.S.; Freire, D.M.G.; Rueda, N.; Fernandez-Lafuente, R. Immobilization of lipases on hydrophobic supports involves the open form of the enzyme. Enzym. Microb. Technol. 2015, 71, 53–57. [Google Scholar] [CrossRef] [PubMed]
  214. dos Santos, J.C.S.; Rueda, N.; Sanchez, A.; Villalonga, R.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Versatility of divinylsulfone supports permits the tuning of CALB properties during its immobilization. RSC Adv. 2015, 5, 35801–35810. [Google Scholar] [CrossRef]
  215. dos Santos, J.C.S.; Rueda, N.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Tuning the catalytic properties of lipases immobilized on divinylsulfone activated agarose by altering its nanoenvironment. Enzym. Microb. Technol. 2015, 77, 1–7. [Google Scholar] [CrossRef]
  216. Dos Santos, J.C.S.; Rueda, N.; Torres, R.; Barbosa, O.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Evaluation of divinylsulfone activated agarose to immobilize lipases and to tune their catalytic properties. Process Biochem. 2015, 50, 918–927. [Google Scholar] [CrossRef]
  217. Tacias-Pascacio, V.G.; Virgen-Ortíz, J.J.; Jiménez-Pérez, M.; Yates, M.; Torrestiana-Sanchez, B.; Rosales-Quintero, A.; Fernandez-Lafuente, R. Evaluation of different lipase biocatalysts in the production of biodiesel from used cooking oil: Critical role of the immobilization support. Fuel 2017, 200, 1–10. [Google Scholar] [CrossRef]
  218. Tacias-Pascacio, V.G.; Peirce, S.; Torrestiana-Sanchez, B.; Yates, M.; Rosales-Quintero, A.; Virgen-Ortíz, J.J.; Fernandez-Lafuente, R. Evaluation of different commercial hydrophobic supports for the immobilization of lipases: Tuning their stability, activity and specificity. RSC Adv. 2016, 6, 100281–100294. [Google Scholar] [CrossRef]
  219. Arana-Peña, S.; Rios, N.S.; Carballares, D.; Mendez-Sanchez, C.; Lokha, Y.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Effects of Enzyme Loading and Immobilization Conditions on the Catalytic Features of Lipase from Pseudomonas fluorescens Immobilized on Octyl-Agarose Beads. Front. Bioeng. Biotechnol. 2020, 8, 36. [Google Scholar] [CrossRef]
  220. Lokha, Y.; Arana-Peña, S.; Rios, N.S.; Mendez-Sanchez, C.; Gonçalves, L.R.B.; Lopez-Gallego, F.; Fernandez-Lafuente, R. Modulating the properties of the lipase from Thermomyces lanuginosus immobilized on octyl agarose beads by altering the immobilization conditions. Enzym. Microb. Technol. 2020, 133, 109461. [Google Scholar] [CrossRef]
  221. Arana-Peña, S.; Rios, N.S.; Carballares, D.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Immobilization of lipases via interfacial activation on hydrophobic supports: Production of biocatalysts libraries by altering the immobilization conditions. Catal. Today 2021, 362, 130–140. [Google Scholar] [CrossRef]
  222. Virgen-Ortíz, J.J.; Tacias-Pascacio, V.G.; Hirata, D.B.; Torrestiana-Sanchez, B.; Rosales-Quintero, A.; Fernandez-Lafuente, R. Relevance of substrates and products on the desorption of lipases physically adsorbed on hydrophobic supports. Enzym. Microb. Technol. 2017, 96, 30–35. [Google Scholar] [CrossRef]
  223. Rueda, N.; Dos Santos, J.C.S.; Torres, R.; Ortiz, C.; Barbosa, O.; Fernandez-Lafuente, R. Improved performance of lipases immobilized on heterofunctional octyl-glyoxyl agarose beads. RSC Adv. 2015, 5, 11212–11222. [Google Scholar] [CrossRef]
  224. Fernandez-Lopez, L.; Pedrero, S.G.; Lopez-Carrobles, N.; Virgen-Ortíz, J.J.; Gorines, B.C.; Otero, C.; Fernandez-Lafuente, R. Physical crosslinking of lipase from Rhizomucor miehei immobilized on octyl agarose via coating with ionic polymers: Avoiding enzyme release from the support. Process Biochem. 2017, 54, 81–88. [Google Scholar] [CrossRef]
  225. Fernandez-Lopez, L.; Virgen-OrtÍz, J.J.; Pedrero, S.G.; Lopez-Carrobles, N.; Gorines, B.C.; Otero, C.; Fernandez-Lafuente, R. Optimization of the coating of octyl-CALB with ionic polymers to improve stability and decrease enzyme leakage. Biocatal. Biotransform. 2018, 36, 47–56. [Google Scholar] [CrossRef]
  226. Pizarro, C.; Brañes, M.C.; Markovits, A.; Fernández-Lorente, G.; Guisán, J.M.; Chamy, R.; Wilson, L. Influence of different immobilization techniques for Candida cylindracea lipase on its stability and fish oil hydrolysis. J. Mol. Catal. B Enzym. 2012, 78, 111–118. [Google Scholar] [CrossRef]
  227. Zaak, H.; Fernandez-Lopez, L.; Otero, C.; Sassi, M.; Fernandez-Lafuente, R. Improved stability of immobilized lipases via modification with polyethylenimine and glutaraldehyde. Enzym. Microb. Technol. 2017, 106, 67–74. [Google Scholar] [CrossRef] [PubMed]
  228. Bi, Y.; Yu, M.; Zhou, H.; Zhou, H.; Wei, P. Biosynthesis of oleyl oleate in solvent-free system by Candida rugosa Lipase (CRL) immobilized in macroporous resin with cross-linking of aldehyde-dextran. J. Mol. Catal. B Enzym. 2016, 133, 1–5. [Google Scholar] [CrossRef]
  229. Fernandez-Lorente, G.; Filice, M.; Lopez-Vela, D.; Pizarro, C.; Wilson, L.; Betancor, L.; Avila, Y.; Guisan, J.M. Cross-Linking of Lipases Adsorbed on Hydrophobic Supports: Highly Selective Hydrolysis of Fish Oil Catalyzed by RML. J. Am. Oil Chem. Soc. 2011, 88, 801–807. [Google Scholar] [CrossRef]
  230. De Almeida, D.G.; Soares Da Silva, R.d.C.F.; Luna, J.M.; Rufino, R.D.; Santos, V.A.; Banat, I.M.; Sarubbo, L.A. Biosurfactants: Promising molecules for petroleum biotechnology advances. Front. Microbiol. 2016, 7, 1718. [Google Scholar] [CrossRef]
  231. Ren, L.; Jia, H.; Yu, M.; Shen, W.; Zhou, H.; Wei, P. Enhanced catalytic ability of Candida rugosa lipase immobilized on pore-enlarged hollow silica microspheres and cross-linked by modified dextran in both aqueous and non-aqueous phases. Biotechnol. Bioprocess Eng. 2013, 18, 888–896. [Google Scholar] [CrossRef]
  232. Bernal, C.; Poveda-Jaramillo, J.C.; Mesa, M. Raising the enzymatic performance of lipase and protease in the synthesis of sugar fatty acid esters, by combined ionic exchange-hydrophobic immobilization process on aminopropyl silica support. Chem. Eng. J. 2018, 334, 760–767. [Google Scholar] [CrossRef]
  233. Vescovi, V.; Kopp, W.; Guisán, J.M.; Giordano, R.L.C.; Mendes, A.A.; Tardioli, P.W. Improved catalytic properties of Candida antarctica lipase B multi-attached on tailor-made hydrophobic silica containing octyl and multifunctional amino-glutaraldehyde spacer arms. Process Biochem. 2016, 51, 2055–2066. [Google Scholar] [CrossRef]
  234. Guajardo, N.; Bernal, C.; Wilson, L.; Cabrera, Z. Selectivity of R-α-monobenzoate glycerol synthesis catalyzed by Candida antarctica lipase B immobilized on heterofunctional supports. Process Biochem. 2015, 50, 1870–1877. [Google Scholar] [CrossRef]
  235. Hirata, D.B.; Albuquerque, T.L.; Rueda, N.; Sánchez-Montero, J.M.; Garcia-Verdugo, E.; Porcar, R.; Fernandez-Lafuente, R. Advantages of Heterofunctional Octyl Supports: Production of 1,2-Dibutyrin by Specific and Selective Hydrolysis of Tributyrin Catalyzed by Immobilized Lipases. ChemistrySelect 2016, 1, 3259–3270. [Google Scholar] [CrossRef]
  236. Souza, P.M.P.; Carballares, D.; Lopez-Carrobles, N.; Gonçalves, L.R.B.; Lopez-Gallego, F.; Rodrigues, S.; Fernandez-Lafuente, R. Enzyme-support interactions and inactivation conditions determine Thermomyces lanuginosus lipase inactivation pathways: Functional and florescence studies. Int. J. Biol. Macromol. 2021, 191, 79–91. [Google Scholar] [CrossRef]
  237. Guimarães, J.R.; Carballares, D.; Rocha-Martin, J.; Alcántara, A.R.; Tardioli, P.W.; Fernandez-Lafuente, R. Heterofunctional Methacrylate Beads Bearing Octadecyl and Vinyl Sulfone Groups: Tricks to Obtain an Interfacially Activated Lipase from Thermomyces lanuginosus and Covalently Attached to the Support. Catalysts 2023, 13, 108. [Google Scholar] [CrossRef]
  238. Hirata, D.B.; Albuquerque, T.L.; Rueda, N.; Virgen-Ortíz, J.J.; Tacias-Pascacio, V.G.; Fernandez-Lafuente, R. Evaluation of different immobilized lipases in transesterification reactions using tributyrin: Advantages of the heterofunctional octyl agarose beads. J. Mol. Catal. B Enzym. 2016, 133, 117–123. [Google Scholar] [CrossRef]
  239. Souza, P.M.P.; Carballares, D.; Gonçalves, L.R.B.; Fernandez-Lafuente, R.; Rodrigues, S. Immobilization of Lipase B from Candida antarctica in Octyl-Vinyl Sulfone Agarose: Effect of the Enzyme-Support Interactions on Enzyme Activity, Specificity, Structure and Inactivation Pathway. Int. J. Mol. Sci. 2022, 23, 14268. [Google Scholar] [CrossRef]
  240. Li, K.; Fan, Y.; He, Y.; Zeng, L.; Han, X.; Yan, Y. Burkholderia cepacia lipase immobilized on heterofunctional magnetic nanoparticles and its application in biodiesel synthesis. Sci. Rep. 2017, 7, 16473. [Google Scholar] [CrossRef]
  241. Bernal, C.; Illanes, A.; Wilson, L. Heterofunctional Hydrophilic–Hydrophobic Porous Silica as Support for Multipoint Covalent Immobilization of Lipases: Application to Lactulose Palmitate Synthesis. Langmuir 2014, 30, 3557–3566. [Google Scholar] [CrossRef] [PubMed]
  242. Rios, N.S.; Mendez-Sanchez, C.; Arana-Peña, S.; Rueda, N.; Ortiz, C.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Immobilization of lipase from Pseudomonas fluorescens on glyoxyl-octyl-agarose beads: Improved stability and reusability. Biochim. Biophys. Acta Proteins Proteom. 2019, 1867, 741–747. [Google Scholar] [CrossRef] [PubMed]
  243. Guajardo, N.; Bernal, C.; Wilson, L.; Cabrera, Z. Asymmetric hydrolysis of dimethyl-3-phenylglutarate in sequential batch reactor operation catalyzed by immobilized Geobacillus thermocatenulatus lipase. Catal. Today 2015, 255, 21–26. [Google Scholar] [CrossRef]
  244. Suescun, A.; Rueda, N.; dos Santos, J.C.S.; Castillo, J.J.; Ortiz, C.; Torres, R.; Barbosa, O.; Fernandez-Lafuente, R. Immobilization of lipases on glyoxyl–octyl supports: Improved stability and reactivation strategies. Process Biochem. 2015, 50, 1211–1217. [Google Scholar] [CrossRef]
  245. Rueda, N.; dos Santos, C.S.; Rodriguez, M.D.; Albuquerque, T.L.; Barbosa, O.; Torres, R.; Ortiz, C.; Fernandez-Lafuente, R. Reversible immobilization of lipases on octyl-glutamic agarose beads: A mixed adsorption that reinforces enzyme immobilization. J. Mol. Catal. B Enzym. 2016, 128, 10–18. [Google Scholar] [CrossRef]
  246. Rueda, N.; Albuquerque, T.L.; Bartolome-Cabrero, R.; Fernandez-Lopez, L.; Torres, R.; Ortiz, C.; Dos Santos, J.C.S.; Barbosa, O.; Fernandez-Lafuente, R. Reversible Immobilization of Lipases on Heterofunctional Octyl-Amino Agarose Beads Prevents Enzyme Desorption. Molecules 2016, 21, 646. [Google Scholar] [CrossRef]
  247. de Albuquerque, T.L.; Rueda, N.; dos Santos, J.C.S.; Barbosa, O.; Ortiz, C.; Binay, B.; Özdemir, E.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Easy stabilization of interfacially activated lipases using heterofunctional divinyl sulfone activated-octyl agarose beads. Modulation of the immobilized enzymes by altering their nanoenvironment. Process Biochem. 2016, 51, 865–874. [Google Scholar] [CrossRef]
  248. Arana-Peña, S.; Mendez-Sanchez, C.; Rios, N.S.; Ortiz, C.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. New applications of glyoxyl-octyl agarose in lipases co-immobilization: Strategies to reuse the most stable lipase. Int. J. Biol. Macromol. 2019, 131, 989–997. [Google Scholar] [CrossRef]
  249. Rios, N.S.; Arana-Peña, S.; Mendez-Sanchez, C.; Ortiz, C.; Gonçalves, L.R.B.; Fernandez-Lafuente, R. Reuse of Lipase from Pseudomonas fluorescens via Its Step-by-Step Coimmobilization on Glyoxyl-Octyl Agarose Beads with Least Stable Lipases. Catalysts 2019, 9, 487. [Google Scholar] [CrossRef]
  250. Carballares, D.; Rocha-Martin, J.; Fernandez-Lafuente, R. Coimmobilization of lipases exhibiting three very different stability ranges. Reuse of the active enzymes and selective discarding of the inactivated ones. Int. J. Biol. Macromol. 2022, 206, 580–590. [Google Scholar] [CrossRef]
  251. Dossat, V.; Combes, D.; Marty, A. Continuous enzymatic transesterification of high oleic sunflower oil in a packed bed reactor: Influence of the glycerol production. Enzym. Microb. Technol. 1999, 25, 194–200. [Google Scholar] [CrossRef]
  252. Séverac, E.; Galy, O.; Turon, F.; Pantel, C.A.; Condoret, J.S.; Monsan, P.; Marty, A. Selection of CalB immobilization method to be used in continuous oil transesterification: Analysis of the economical impact. Enzym. Microb. Technol. 2011, 48, 61–70. [Google Scholar] [CrossRef]
  253. Lopresto, G.C.; Calabrò, V.; Woodley, J.M.; Tufvesson, P. Enzymatic kinetic study on the enzymatic esterification of octanoic acid and hexanol by immobilized Candida antarctica lipase B. J. Mol. Catal. B Enzym. 2014, 110, 64–71. [Google Scholar] [CrossRef]
  254. Varma, M.N.; Madras, G. Effect of chain length of alcohol on the lipase-catalyzed esterification of propionic acid in supercritical carbon dioxide. Appl. Biochem. Biotechnol. 2010, 160, 2342–2354. [Google Scholar] [CrossRef]
  255. Alves, J.S.; Garcia-Galan, C.; Schein, M.F.; Silva, A.M.; Barbosa, O.; Ayub, M.A.Z.; Fernandez-Lafuente, R.; Rodrigues, R.C. Combined effects of ultrasound and immobilization protocol on butyl acetate synthesis catalyzed by CALB. Molecules 2014, 19, 9562–9576. [Google Scholar] [CrossRef] [PubMed]
  256. Martins, A.B.; Friedrich, J.L.R.; Cavalheiro, J.C.; Garcia-Galan, C.; Barbosa, O.; Ayub, M.A.Z.; Fernandez-Lafuente, R.; Rodrigues, R.C. Improved production of butyl butyrate with lipase from Thermomyces lanuginosus immobilized on styrene-divinylbenzene beads. Bioresour. Technol. 2013, 134, 417–422. [Google Scholar] [CrossRef] [PubMed]
  257. Poppe, J.K.; Garcia-Galan, C.; Matte, C.R.; Fernandez-Lafuente, R.; Rodrigues, R.C.; Ayub, M.A.Z. Optimization of synthesis of fatty acid methyl esters catalyzed by lipase B from Candida antarctica immobilized on hydrophobic supports. J. Mol. Catal. B Enzym. 2013, 94, 51–56. [Google Scholar] [CrossRef]
  258. Fallavena, L.P.; Antunes, F.H.F.; Alves, J.S.; Paludo, N.; Ayub, M.A.Z.; Fernandez-Lafuente, R.; Rodrigues, R.C. Ultrasound technology and molecular sieves improve the thermodynamically controlled esterification of butyric acid mediated by immobilized lipase from Rhizomucor miehei. RSC Adv. 2014, 4, 8675–8681. [Google Scholar] [CrossRef]
  259. Martins, A.B.; Schein, M.F.; Friedrich, J.L.R.; Fernandez-Lafuente, R.; Ayub, M.A.Z.; Rodrigues, R.C. Ultrasound-assisted butyl acetate synthesis catalyzed by Novozym 435: Enhanced activity and operational stability. Ultrason. Sonochem. 2013, 20, 1155–1160. [Google Scholar] [CrossRef]
  260. Paludo, N.; Alves, J.S.; Altmann, C.; Ayub, M.A.Z.; Fernandez-Lafuente, R.; Rodrigues, R.C. The combined use of ultrasound and molecular sieves improves the synthesis of ethyl butyrate catalyzed by immobilized Thermomyces lanuginosus lipase. Ultrason. Sonochem. 2014, 22, 89–94. [Google Scholar] [CrossRef]
  261. Alves, J.S.; Garcia-Galan, C.; Danelli, D.; Paludo, N.; Barbosa, O.; Rodrigues, R.C.; Fernandez-Lafuente, R. Use of Lecitase-Ultra immobilized on styrene-divinylbenzene beads as catalyst of esterification reactions: Effects of ultrasounds. Catal. Today 2015, 255, 27–32. [Google Scholar] [CrossRef]
  262. Poppe, J.K.; Matte, C.R.; Fernandez-Lafuente, R.; Rodrigues, R.C.; Ayub, M.A.Z. Transesterification of Waste Frying Oil and Soybean Oil by Combi-lipases Under Ultrasound-Assisted Reactions. Appl. Biochem. Biotechnol. 2018, 186, 576–589. [Google Scholar] [CrossRef] [PubMed]
  263. José, C.; Bonetto, R.D.; Gambaro, L.A.; Torres, M.P.G.; Foresti, M.L.; Ferreira, M.L.; Briand, L.E. Enzymatic Investigation of the causes of deactivation–degradation of the commercial biocatalyst Novozym® 435 in ethanol and ethanol–aqueous media. J. Mol. Catal. B Enzym. 2011, 71, 95–107. [Google Scholar] [CrossRef]
  264. Baek, Y.; Lee, J.; Son, J.; Lee, T.; Sobhan, A.; Lee, J.; Koo, S.M.; Shin, W.H.; Oh, J.M.; Park, C. Enzymatic synthesis of formate ester through immobilized lipase and its reuse. Polymers 2020, 12, 1802. [Google Scholar] [CrossRef]
  265. Seo, J.; Shin, M.; Lee, J.; Lee, T.; Oh, J.M.; Park, C. Novel and highly efficient lipase-catalyzed esterification of formic acid with hexanol for waste gas reutilization. J. Ind. Eng. Chem. 2021, 93, 430–435. [Google Scholar] [CrossRef]
  266. Martins, A.B.; Graebin, N.G.; Lorenzoni, A.S.G.; Fernandez-Lafuente, R.; Ayub, M.A.Z.; Rodrigues, R.C. Rapid and high yields of synthesis of butyl acetate catalyzed by Novozym 435: Reaction optimization by response surface methodology. Process Biochem. 2011, 46, 2311–2316. [Google Scholar] [CrossRef]
  267. Hari Krishna, S.; Divakar, S.; Prapulla, S.G.; Karanth, N.G. Enzymatic synthesis of isoamyl acetate using immobilized lipase from Rhizomucor miehei. J. Biotechnol. 2001, 87, 193–201. [Google Scholar] [CrossRef]
  268. Güvenç, A.; Kapucu, N.; Kapucu, H.; Aydoǧan, Ö.; Mehmetoǧlu, Ü. Enzymatic esterification of isoamyl alcohol obtained from fusel oil: Optimization by response surface methodolgy. Enzym. Microb. Technol. 2007, 40, 778–785. [Google Scholar] [CrossRef]
  269. Nyari, N.; Paulazzi, A.; Zamadei, R.; Steffens, C.; Zabot, G.L.; Tres, M.V.; Zeni, J.; Venquiaruto, L.; Dallago, R.M. Synthesis of isoamyl acetate by ultrasonic system using Candida antarctica lipase B immobilized in polyurethane. J. Food Process Eng. 2018, 41, e12812. [Google Scholar] [CrossRef]
  270. Narwal, S.K.; Saun, N.K.; Dogra, P.; Gupta, R. Green synthesis of isoamyl acetate via silica immobilized novel thermophilic lipase from Bacillus aerius. Russ. J. Bioorg. Chem. 2016, 42, 69–73. [Google Scholar] [CrossRef]
  271. De Oliveira, T.P.; Pereira, M.; Santos, F. Incorporation of metallic particles in activated carbon used in lipase immobilization for production of isoamyl acetate. J. Chem. Technol. Biotechnol. 2022, 97, 1736–1746. [Google Scholar] [CrossRef]
  272. Zhang, Y.; Feng, T.; Ni, X.; Xia, J.; Suo, H.; Yan, L.; Zou, B. Immobilized lipase based on SBA-15 adsorption and gel embedding for catalytic synthesis of isoamyl acetate. Food Biosci. 2024, 60, 104427. [Google Scholar] [CrossRef]
  273. Diaz, M.D.R.; Gómez, J.M.; Díaz-Suelto, B.; García-Sanz, A. Enzymatic synthesis of short-chain esters in n-hexane and supercritical carbon dioxide: Effect of the acid chain length. Eng. Life Sci. 2010, 10, 171–176. [Google Scholar] [CrossRef]
  274. Bourg-Garros, S.; Razafindramboa, N.; Pavia, A.A. Optimization of lipase-catalyzed synthesis of (z)-3-hexen-1-yl acetate by direct esterification in hexane and a solvent-free medium. Enzym. Microb. Technol. 1998, 22, 240–245. [Google Scholar] [CrossRef]
  275. Dai, W.C.; Chiu, S.J.; Huang, D.Y.; Juan, H.Y.; Chen, C.Y.; Chen, S.S.; Su, C.H.; Li, S.Y. Lipase-catalyzed synthesis of butyl propionate in solvent-free system: Optimization by response surface methodology. J. Taiwan Inst. Chem. Eng. 2014, 45, 2233–2237. [Google Scholar] [CrossRef]
  276. Bhavsar, K.V.; Yadav, G.D. Microwave assisted solvent-free synthesis of n-butyl propionate by immobilized lipase as catalyst. Biocatal. Agric. Biotechnol. 2018, 14, 264–269. [Google Scholar] [CrossRef]
  277. Jaiswal, K.S.; Rathod, V.K. Acoustic cavitation promoted lipase catalysed synthesis of isobutyl propionate in solvent free system: Optimization and kinetic studies. Ultrason. Sonochem. 2018, 40, 727–735. [Google Scholar] [CrossRef]
  278. Pires-Cabral, P.; da Fonseca, M.M.R.; Ferreira-Dias, S. Esterification activity and operational stability of Candida rugosa lipase immobilized in polyurethane foams in the production of ethyl butyrate. Biochem. Eng. J. 2010, 48, 246–252. [Google Scholar] [CrossRef]
  279. Rodriguez-Nogales, J.M.; Roura, E.; Contreras, E. Biosynthesis of ethyl butyrate using immobilized lipase: A statistical approach. Process Biochem. 2005, 40, 63–68. [Google Scholar] [CrossRef]
  280. Rahman, I.N.A.; Manan, F.M.A.; Marzuki, N.H.C.; Mahat, N.A.; Attan, N.; Keyon, A.S.A.; Jamalis, J.; Aboul-Enein, H.Y.; Wahab, R.A. A statistical approach for optimizing the high yield green production of the flavor ester butyl butyrate. J. Teknol. 2017, 79, 141–151. [Google Scholar] [CrossRef]
  281. Sjöblom, M.; Risberg, P.; Filippova, A.; Öhrman, O.G.W.; Rova, U.; Christakopoulos, P. In Situ Biocatalytic Synthesis of Butyl Butyrate in Diesel and Engine Evaluations. ChemCatChem 2017, 9, 4529–4537. [Google Scholar] [CrossRef]
  282. Costa-Silva, T.A.; Carvalho, A.K.F.; Souza, C.R.F.; De Castro, H.F.; Bachmann, L.; Said, S.; Oliveira, W.P. General Enhancement lipase activity via immobilization onto chitosan beads used as seed particles during fluidized bed drying: Application in butyl butyrate production. Appl. Catal. A Gen. 2021, 622, 118217. [Google Scholar] [CrossRef]
  283. Sharma, S.; Kumar, S.; Sahney, R.; Dahiya, P. Green synthesis of iron-alginate nanoparticle for Taguchi-assisted immobilization of Candida rugosa lipase and its application in the synthesis of butyl butyrate ester. Process Biochem. 2024, 139, 81–92. [Google Scholar] [CrossRef]
  284. Gonçalves, G.R.F.; Reinert, O.; Gandolfi, R.; Junqueira, M.; Brito, P.; Cristina, R.; Bonomo, F.; Ilh, C.; Veloso, C.M. Immobilization of porcine pancreatic lipase on activated carbon by adsorption and covalent bonding and its application in the synthesis of butyl butyrate. Process Biochem. 2021, 111, 114–123. [Google Scholar] [CrossRef]
  285. Kaur, P.; Kumar, A. Amino functionalization of magnetic multiwalled carbon nanotubes with flexible hydrophobic spacer for immobilization of Candida rugosa lipase and application in biocatalytic production of fruit flavour esters ethyl butyrate and butyl butyrate. Appl. Nanosci. 2023, 13, 4291–4311. [Google Scholar] [CrossRef]
  286. Anschau, A.; da Mota Huerta, K.; Rêgo, T.V.; de Oliveira, J.M.G.; Borba, C.M.; Kalil, S.J.; Burkert, C.A.V.; de Medeiros Burkert, J.F. Enzymatic synthesis optimization of isoamyl butyrate from fusel oil. Acta Sci. Biol. Sci. 2021, 43, e54966. [Google Scholar] [CrossRef]
  287. Hari Krishna, S.; Sattur, A.P.; Karanth, N.G. Lipase-catalyzed synthesis of isoamyl isobutyrate—Optimization using a central composite rotatable design. Process Biochem. 2001, 37, 9–16. [Google Scholar] [CrossRef]
  288. Santos, M.M.O.; Menezes, L.H.S.; Santo, E.L.E.; Carvalho, M.S.; Gonçalves, M.S.; Tavares, I.M.C.; Mendes, A.A.; Ruiz, H.A.; Salay, L.C.; Franco, M.; et al. Synthesis of hexyl butyrate (apple and citrus aroma) by Candida rugosa lipase immobilized on Diaion HP-20 using the Box-Behnken design. Food Sci. Biotechnol. 2023, 32, 689–696. [Google Scholar] [CrossRef]
  289. Raghavendra, T.; Sayania, D.; Madamwar, D. Synthesis of the ‘green apple ester’ ethyl valerate in organic solvents by Candida rugosa lipase immobilized in MBGs in organic solvents: Effects of immobilization and reaction parameters. J. Mol. Catal. B Enzym. 2010, 63, 31–38. [Google Scholar] [CrossRef]
  290. Bhavsar, K.V.; Yadav, G.D. Process intensification by microwave irradiation in immobilized-lipase catalysis in solvent-free synthesis of ethyl valerate. Mol. Catal. 2018, 461, 34–39. [Google Scholar] [CrossRef]
  291. Cebrián-García, S.; Balu, A.M.; García, A.; Luque, R. Sol-gel immobilisation of lipases: Towards active and stable biocatalysts for the esterification of valeric acid. Molecules 2018, 23, 2283. [Google Scholar] [CrossRef] [PubMed]
  292. Hussin, N.H.; Wahab, R.A.; Elias, N.; Jacob, A.G.; Zainal-Abidin, M.H.; Abdullah, F.; Sulaiman, N.J.; Misson, M. Electrospun magnetic nanocellulose–polyethersulfone-conjugated Aspergillus oryzae lipase for synthesis of ethyl valerate. Membranes 2021, 11, 972. [Google Scholar] [CrossRef]
  293. Padilha, G.S.; De Barros, M.; Alegre, R.M.; Tambourgi, E.B. Production of ethyl valerate from Burkholderia cepacia lipase immobilized in alginate. Chem. Eng. Trans. 2013, 32, 1063–1068. [Google Scholar] [CrossRef]
  294. Jacob, A.G.; Wahab, R.A.; Chandren, S.; Jumbri, K.; Mahmood, W.M.A. Physicochemical properties and operational stability of Taguchi design-optimized Candida rugosa lipase supported on biogenic silica/magnetite/graphene oxide for ethyl valerate synthesis. Adv. Powder Technol. 2022, 33, 103374. [Google Scholar] [CrossRef]
  295. Elias, N.; Wahab, R.A.; Chandren, S.; Jamalis, J.; Mahat, N.A.; Jye, L.W. Structure and properties of lipase activated by cellulose-silica polyethersulfone membrane for production of pentyl valerate. Carbohydr. Polym. 2020, 245, 116549. [Google Scholar] [CrossRef]
  296. Musa, N.; Latip, W.; Abd Rahman, R.N.Z.; Salleh, A.B.; Ali, M.S.M. Immobilization of an antarctic Pseudomonas AMS8 lipase for low temperature ethyl hexanoate synthesis. Catalysts 2018, 8, 234. [Google Scholar] [CrossRef]
  297. Shin, M.; Seo, J.; Baek, Y.; Lee, T.; Jang, M.; Park, C. Novel and efficient synthesis of phenethyl formate via enzymatic esterification of formic acid. Biomolecules 2020, 10, 70. [Google Scholar] [CrossRef] [PubMed]
  298. Sá, A.G.A.; de Meneses, A.C.; Lerin, L.A.; de Araújo, P.H.H.; Sayer, C.; de Oliveira, D. Biocatalysis of aromatic benzyl-propionate ester by different immobilized lipases. Bioprocess Biosyst. Eng. 2018, 41, 585–591. [Google Scholar] [CrossRef]
  299. de Meneses, A.C.; Lerin, L.A.; Araújo, P.H.H.; Sayer, C.; de Oliveira, D. Benzyl propionate synthesis by fed-batch esterification using commercial immobilized and lyophilized CALB lipase. Bioprocess Biosyst. Eng. 2019, 42, 1625–1634. [Google Scholar] [CrossRef]
  300. Jeromin, G.E.; Zoor, A. A new irreversible enzyme-aided esterification method in organic solvents. Biotechnol. Lett. 2008, 30, 925–928. [Google Scholar] [CrossRef]
  301. Chen, J.; Zhang, Y.; Zhong, H.; Zhu, H.; Wang, H.; Goh, K.L.; Zhang, J.; Zheng, M. Efficient and sustainable preparation of cinnamic acid flavor esters by immobilized lipase microarray. LWT 2023, 173, 114322. [Google Scholar] [CrossRef]
  302. Zaks, A.; Klibanov, A.M. Enzymatic catalysis in organic media at 100 °C. Science 1984, 224, 1249–1251. [Google Scholar] [CrossRef]
  303. Klibanov, A.M. Improving enzymes by using them in organic solvents. Nature 2001, 409, 241–246. [Google Scholar] [CrossRef]
  304. Valivety, R.H.; Johnston, G.A.; Suckling, C.J.; Halling, P.J. Solvent effects on biocatalysis in organic systems: Equilibrium position and rates of lipase catalyzed esterification. Biotechnol. Bioeng. 1991, 38, 1137–1143. [Google Scholar] [CrossRef]
  305. García-Alles, L.F.; Gotor, V. Alcohol Inhibition and Specificity Studies of Lipase B from Candida antarctica in Organic Solvents. Biotechnol. Bioeng. 1998, 59, 163–170. [Google Scholar] [CrossRef]
  306. Khmelnitsky, Y.L.; Levashov, A.V.; Klyachko, N.L.; Martinek, K. Engineering biocatalytic systems in organic media with low water content. Enzym. Microb. Technol. 1988, 10, 710–724. [Google Scholar] [CrossRef]
  307. Branco, R.J.F.; Graber, M.; Denis, V.; Pleiss, J. Molecular mechanism of the hydration of Candida antarctica lipase B in the gas phase: Water adsorption isotherms and molecular dynamics simulations. ChemBioChem 2009, 10, 2913–2919. [Google Scholar] [CrossRef]
  308. Trodler, P.; Pleiss, J. Modeling structure and flexibility of Candida antarctica lipase B in organic solvents. BMC Struct. Biol. 2008, 8, 9. [Google Scholar] [CrossRef] [PubMed]
  309. Rehm, S.; Trodler, P.; Pleiss, J. Solvent-induced lid opening in lipases: A molecular dynamics study. Protein Sci. 2010, 19, 2122–2130. [Google Scholar] [CrossRef]
  310. Cebrián-García, S.; Balu, A.M.; Luque, R. Ultrasound-assisted esterification of valeric acid to alkyl valerates promoted by biosilicified lipases. Front. Chem. 2018, 6, 197. [Google Scholar] [CrossRef]
  311. Ewing, T.A.; Nouse, N.; van Lint, M.; van Haveren, J.; Hugenholtz, J.; van Es, D.S. Fermentation for the production of biobased chemicals in a circular economy: A perspective for the period 2022–2050. Green Chem. 2022, 24, 6373–6405. [Google Scholar] [CrossRef]
  312. Kovalenko, G.; Perminova, L.; Pykhtina, M.; Beklemishev, A. Lipase-active heterogeneous biocatalysts for enzymatic synthesis of short-chain aroma esters. Biocatal. Agric. Biotechnol. 2021, 36, 102124. [Google Scholar] [CrossRef]
  313. Kütt, A.; Selberg, S.; Kaljurand, I.; Tshepelevitsh, S.; Heering, A.; Darnell, A.; Kaupmees, K.; Piirsalu, M.; Leito, I. pKa values in organic chemistry—Making maximum use of the available data. Tetrahedron Lett. 2018, 59, 3738–3748. [Google Scholar] [CrossRef]
  314. Rosell, C.M.; Terreni, M.; Fernandez-Lafuente, R.; Guisan, J.M. A criterion for the selection of monophasic solvents for enzymatic synthesis. Enzym. Microb. Technol. 1998, 23, 64–69. [Google Scholar] [CrossRef]
  315. van Schie, M.M.C.H.; Spöring, J.-D.; Bocola, M.; Domínguez de María, P.; Rother, D. Applied biocatalysis beyond just buffers—From aqueous to unconventional media. Options and guidelines. Green Chem. 2021, 23, 3191–3206. [Google Scholar] [CrossRef] [PubMed]
  316. Dai, Y.; Van Spronsen, J.; Witkamp, G. Natural deep eutectic solvents as new potential media for green technology. Anal. Chim. Acta 2013, 766, 61–68. [Google Scholar] [CrossRef]
  317. Sheldon, R.A.; Bode, M.L.; Mathebula, N. Green and sustainable solvents for biocatalytic oxidations. Curr. Opin. Green Sustain. Chem. 2023, 39, 100741. [Google Scholar] [CrossRef]
  318. Hari Krishna, S.; Karanth, N.G. Lipase-catalyzed synthesis of isoamyl butyrate: A kinetic study. Biochim. Biophys. Acta—Protein Struct. Mol. Enzymol. 2001, 1547, 262–267. [Google Scholar] [CrossRef]
  319. Horikoshi, S.; Schiffmann, R.F.; Fukushima, J.; Serpone, N. Microwave Chemical and Materials Processing: A Tutorial; Springer: Berlin/Heidelberg, Germany, 2017; pp. 1–393. [Google Scholar] [CrossRef]
  320. Santos, D.; da Rocha, E.C.L.; Santos, R.L.M.; Cancelas, A.J.; Franceschi, E.; Santos, A.F.; Fortuny, M.; Dariva, C. Demulsification of water-in-crude oil emulsions using single mode and multimode microwave irradiation. Sep. Purif. Technol. 2017, 189, 347–356. [Google Scholar] [CrossRef]
  321. Chandrasekaran, S.; Ramanathan, S.; Basak, T. Microwave material processing—A review. AIChE J. 2012, 58, 330–363. [Google Scholar] [CrossRef]
  322. Leadbeater, N.E.; Stencel, L.M. Fast, Easy Preparation of Biodiesel Using Microwave Heating. Energy Fuels 2006, 20, 2281–2283. [Google Scholar] [CrossRef]
  323. Choedkiatsakul, I.; Ngaosuwan, K.; Assabumrungrat, S.; Mantegna, S.; Cravotto, G. Biodiesel production in a novel continuous flow microwave reactor. Renew. Energy 2015, 83, 25–29. [Google Scholar] [CrossRef]
  324. Yu, D.; Tian, L.; Ma, D.; Wu, H.; Wang, Z.; Wang, L.; Fang, X. Microwave-assisted fatty acid methyl ester production from soybean oil by Novozym 435. Green Chem. 2010, 12, 844–850. [Google Scholar] [CrossRef]
  325. Khan, N.R.; Rathod, V.K. Microwave assisted enzymatic synthesis of speciality esters: A mini-review. Process Biochem. 2018, 75, 89–98. [Google Scholar] [CrossRef]
  326. Hasegawa, S.; Azuma, M.; Takahashi, K. Enzymatic esterification of lactic acid, utilizing the basicity of particular polar organic solvents to suppress the acidity. J. Chem. Technol. Biotechnol. Int. Res. Process Environ. Clean Technol. 2008, 83, 1503–1510. [Google Scholar] [CrossRef]
  327. Laszlo, J.A.; Jackson, M.; Blanco, R.M. Active-site titration analysis of surface influences on immobilized Candida antarctica lipase B activity. J. Mol. Catal. B Enzym. 2011, 69, 60–65. [Google Scholar] [CrossRef]
  328. Chen, B.; Hu, J.; Miller, E.M.; Xie, W.; Cai, M.; Gross, R.A. Candida antarctica Lipase B Chemically Immobilized on Epoxy-Activated Micro- and Nanobeads: Catalysts for Polyester Synthesis. Biomacromolecules 2008, 9, 463–471. [Google Scholar] [CrossRef]
  329. Chen, B.; Miller, E.M.; Miller, L.; Maikner, J.J.; Gross, R.A. Effects of Macroporous Resin Size on Candida antarctica Lipase B Adsorption, Fraction of Active Molecules, and Catalytic Activity for Polyester Synthesis. Langmuir 2007, 23, 1381–1387. [Google Scholar] [CrossRef]
  330. Tambe, A.; Vyasarayani, R.; Datla, A.; Ponrathnam, S.; Demnerova, K. Macroporous Poly (Vinyl Acetate-Co-Divinyl Benzene) Copolymer Beads as Macroporous Poly (Vinyl Acetate-Co-Divinyl Benzene) Copolymer Beads as Adsorptive Support for the Direct Immobilization of Candida antarctica Lipase B. Enzym. Eng. 2015, 4, 1000130. [Google Scholar]
  331. Nicolás, P.; Lassalle, V.L.; Ferreira, M.L. About the role of typical spacer/crosslinker on the design of efficient magnetic biocatalysts based on nanosized magnetite. J. Mol. Catal. B Enzym. 2015, 122, 296–304. [Google Scholar] [CrossRef]
  332. Weinberg, S.; Pellis, A.; Comerford, J.W.; Farmer, T.J.; Guebitz, G.M. Efficient Physisorption of Candida Antarctica Lipase B on Polypropylene Beads and Application for Polyester Synthesis. Catalysts 2018, 8, 369. [Google Scholar] [CrossRef]
  333. Zhao, H.; Song, Z. Migration of reactive trace compounds from Novozym® 435 into organic solvents and ionic liquids. Biochem. Eng. J. 2010, 49, 113–118. [Google Scholar] [CrossRef]
  334. Pires-cabral, P.; Dubreucq, E.; Fonseca, M.M.R.; Ferreira-dias, S. Partitioning of water in organic systems with lipase immobilized in polyurethane foams. Biochem. Eng. J. 2005, 26, 29–37. [Google Scholar] [CrossRef]
  335. Ognjanovic, N.; Bezbradica, D.; Knezevic-Jugovic, Z. Enzymatic conversion of sunflower oil to biodiesel in a solvent-free system: Process optimization and the immobilized system stability. Bioresour. Technol. 2009, 100, 5146–5154. [Google Scholar] [CrossRef] [PubMed]
  336. Foresti, M.L.; Galle, M.; Ferreira, M.L.; Briand, L.E. Enantioselective esterification of ibuprofen with ethanol as reactant and solvent catalyzed by immobilized lipase: Experimental and molecular modeling aspects. J. Chem. Technol. Biotechnol. 2009, 84, 1461–1473. [Google Scholar] [CrossRef]
  337. Toledo, M.V.; José, C.; Collins, S.E.; Bonetto, R.D.; Ferreira, M.L.; Briand, L.E. Enzymatic Esterification of R/S-ketoprofen with 2-propanol as reactant and solvent catalyzed by Novozym® 435 at selected conditions. J. Mol. Catal. B Enzym. 2012, 83, 108–119. [Google Scholar] [CrossRef]
  338. Ru, M.T.; Dordick, J.S.; Reimer, J.A.; Clark, D.S. Optimizing the salt-induced activation of enzymes in organic solvents: Effects of lyophilization time and water content. Biotechnol. Bioeng. 1999, 63, 233–241. [Google Scholar] [CrossRef]
  339. Leszczak, J.-P.; Tran-Minh, C. Optimized enzymatic synthesis of methyl benzoate in organic medium. Operating conditions and impact of different factors on kinetics. Biotechnol. Bioeng. 1998, 60, 356–361. [Google Scholar] [CrossRef]
  340. Sun, S.Y.; Xu, Y.; Wang, D. Novel minor lipase from Rhizopus chinensis during solid-state fermentation: Biochemical characterization and its esterification potential for ester synthesis. Bioresour. Technol. 2009, 100, 2607–2612. [Google Scholar] [CrossRef]
  341. Cybulski, A.; Stankiewicz, A.; Edvinsson Albers, R.K.; Moulijn, J.A. Monolithic reactors for fine chemicals industries: A comparative analysis of a monolithic reactor and a mechanically agitated slurry reactor. Chem. Eng. Sci. 1999, 54, 2351–2358. [Google Scholar] [CrossRef]
  342. Hagen, L.H.; Vivekanand, V.; Linjordet, R.; Pope, P.B.; Eijsink, V.G.H.; Horn, S.J. Microbial community structure and dynamics during co-digestion of whey permeate and cow manure in continuous stirred tank reactor systems. Bioresour. Technol. 2014, 171, 350–359. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The chemical equilibrium of an esterification reaction with the formation of a second phase.
Figure 1. The chemical equilibrium of an esterification reaction with the formation of a second phase.
Catalysts 15 00375 g001
Table 1. Short-chain ester syntheses from aliphatic reagents using immobilized lipases.
Table 1. Short-chain ester syntheses from aliphatic reagents using immobilized lipases.
EsterBiocatalyst
Biocat. Loading
Reaction ConditionsAcid:
Alcohol Molar Ratio
SolventConversion (%)Concentration (g L−1)Productivity
(g L−1 h−1)
Ref.
Octyl FormateNovozym® 435
15 g L−1
Orbital shaker, 150 rpm, 1 h, 40 °C1:71,2-dichloro-
ethane
96.57.67.6[264]
Hexyl FormateNovozym® 435
15 g L−1
Orbital shaker, 150 rpm, 1.5 h, 40 °C1:51,2-dichloro-
ethane
98.312.88.5[265]
Ethyl AcetateNovozym® 435
8 g L−1
Orbital shaker, 150 rpm, 40 °C1:4.4n-heptane
solvent-free
Not ment-
ioned (only reaction rate)
--[13]
Ethyl AcetateNovozym® 435
6% (w/v)
Ultrasound (150 W/28 kHz), ~3 h1:4n-hexane~82.0--[24]
Butyl AcetateLipase B from Candida antarctica immobilized onto porous styrene-divinylben-zene beads 10.0% (w/w)
Novozym® 435
10.0% (w/w)
Orbital shaker, 200 rpm, 2 h
Orbital shaker, 200 rpm, 2 h
1:5
1:3
n-hexane
n-hexane
~90.0
~90.0
31.4
31.4
15.7
15.7
[182]
Butyl AcetateCandida antarctica immobilized onto porous styrene-divinylben-zene beads
7.5% (w/w)
Ultrasound (220 W/40 kHz), ~49 °C, 1.5 h1:3.46n-hexane90.031.420.9[255]
Butyl AcetateRhizopus oryzae lipase immobilized onto Celite 545
500 U in 5 mL
Orbital shaker, 200 rpm, 37 °C, ~12 h [a]1:1solvent-free
n-heptane
n-hexane
60.0
80.0
76.7
334.0
18.6
17.6
27.8
1.6
1.5
[203]
Butyl AcetateNovozym® 435
7.5% (w/w)
Orbital shaker, 200 rpm, 2.5 h, 40 °C1:3n-hexane~90.031.415.7[266]
Butyl AcetateNovozym® 435
7.0% (w/w)
Ultrasound, 2.5 h, 46 °C1:3.6n-hexane94.032.813.1[260]
Isopentyl AcetateLipozyme® RM IM
9.7% (w/w)
Orbital shaker, 150 rpm, 72 h, 40 °C1:4n-heptane95.31502.1[267]
Pentyl AcetateLipozyme® 435
10.0% (w/w acid)
Shaken glass reactor, 100 rpm, 8 h, 40 °C1:2solvent-free89.0398.349.8[21]
Isopentyl AcetateNovozym® 435
5.0% (w/w)
Orbital shaker, 150 rpm, 6 h, 30 °C1:2solvent-free80.00.57 g g−1-[268]
Isopentyl AcetateStaphylococcus simulans lipase immobilized onto CaCO3
60 U
Screw-capped tubes, 200 rpm, 8 h, 37 °C [a]1:2solvent-free64.0495.761.2[61]
Isopentyl AcetateCandida antarctica immobilized onto polyurethane
10% (w/w)
Mechanical agitation, 160 rpm, 6 h, 64 °C
Ultrasonic power, 105 W, 1 h, 65 °C
1:7
1:9
solvent-free94.4
95.4
158.2
125.7
26.3
125.7
[269]
Isopentyl AcetateBacillus aerus lipase immobilized onto silica gel cross-linked with glutaraldehyde
1% (w/w)
Constant shaking, 10 h, 55 °C1:1solvent-free68.065.10.7[270]
Isopentyl AcetatePorcine pancreas lipase immobilized onto activated carbon
0.5 g
Orbital shaker, 200 rpm, 4 h, 40 °C1:1n-hexane93.029.17.28[271]
Isopentyl AcetateCandida rugosa lipase immobilized onto silica support/sodium alginate/calcium chloride
4 mg protein
Orbital shaker, 200 rpm, 8 h, 50 °C1:2.6n-hexane85.2374.046.7[272]
Hexyl AcetateNovozym® 435
13.8 g mol−1
Orbital shaker, 200 rpm, 3 h, 40 °C1:1n-hexane93.067.122.3[273]
3-Hexen-1-yl AcetateNovozym® 435
2% (w/w)
9% (w/w)
Round bottom flask, 250 rpm, ~8 h, 40 °C
Round bottom flask, 250 rpm, 24 h, 40 °C
1:1n-hexane
solvent-free
94.0
70.0
100.2
566.5
12.5
23.6
[274]
Ethyl PropionateBacillus coagulants (MTCC-6375) lipase immobilized onto polyhydrogel 31.4% (w/w)Water-batch-incub-
ator shaker and molecular sieves, 160 rpm, 9 h, 65 °C
1:3n-heptane89.39.11.0[197]
Ethyl PropionatePseudomonas aeruginosa (BTS-2) lipase immobilized onto polyhydrogel network 23.6% (w/w)Orbital shaker, 160 rpm, 3 h, 65 °C1:3n-nonane94.09.63.2[53]
Butyl PropionateNovozym® 435
2.35% (w/w acid)
Orbital shaker, 130 rpm, ~25 h, ~43 °C1:9solvent-free93.7132.55.3[275]
Butyl PropionateNovozym® 435
Lipozyme® RM IM
Lipozyme® TL-IM
0.98% (w/w) each
Microwave, 50 W, 400 rpm, 8 h, 60 °C1:12solvent-free92.0
14.0
9.0
102.2
15.6
10.0
12.8
1.9
1.1
[276]
Isobutyl PropionateFermase® CALB 1000
4% (w/w)
Ultrasound, 40 W, 25 kHz, 150 rpm, 3 h, 60 °C1:3solvent-free95.1354.5118.1[277]
Isobutyl PropionateNovozym® 435
5% (w/w)
Orbital shaker, 300 rpm, 10 h, 40 °C1:3solvent-free92.5345.034.5[37]
Hexyl PropionateNovozym® 435
13.8 g mol−1
Orbital shaker, 200 rpm, 3 h, 40 °C1:1n-hexane94.074.324.7[273]
Ethyl ButyrateCandida rugosa lipase immobilized onto cotton cloth
25.4% (w/w)
Orbital shaker, 220 rpm, 24 h, 25 °C2.4:1Cyclohexane91.226.51.1[207]
Ethyl ButyrateCandida rugosa lipase immobilized into polyurethane foams
1560 mg
Continuous packed-bed reactor, 260 min, residence time, 36 h, 30 °C1:1.5n-hexane80.732.70.7[278]
Ethyl ButyrateNovozym® 435
7% (w/w)
Orbital shaker, 150 rpm, 96 h, 34 °C1:3n-heptane72.933.90.4[279]
Ethyl ButyrateCandida antarctica lipase B immobilized onto styrene beads
10% (w/w)
Novozym® 435
7.5% (w/w)
Orbital shaker, 200 rpm, 1 h, 37.5 °C [a]
Orbital shaker, 200 rpm, 1 h, 44 °C [a]
1:4n-hexane
n-hexane
86.2
87.7
69.9
70.7
69.9
70.7
[183]
Isobutyl ButyrateLipozyme® RM-IM
14% (w/w)
Ultrasound, 2.5 h, 45 °C1:1n-hexane86.0--[259]
Butyl ButyrateNovozym® 435
-
Shaken round bottom flask, ~10 h, 56 °C1:3.93solvent-free99.6--[280]
Butyl ButyrateThermomyces lanuginosus lipase immobilized onto glyoxyl-agarose beads
10% (w/w)
Orbital shaker, 200 rpm, 24 h, 37 °C1:1heptane85.0--[194]
Butyl ButyrateNovozym® 435
12% (w/w)
Round-bottom flasks, 250 rpm, 0.5 h, 57 °C1:1diesel oil90.0129.8259.6[281]
Butyl ButyrateCercospora kikuchii lipase immobilized onto chitosan beads
42.0% (w/w)
45 °C1:1.5n-heptane95.018.03.0[282]
Butyl ButyrateCandida rugosa lipase immobilized onto iron-alginate nanoparticles
-
Orbital shaker, 100 rpm, 6 h, 50 °C1:4solvent-free90.0--[283]
Butyl ButyratePorcine pancreas lipase immobilized onto activated carbon
0.5 g
Orbital shaker, 200 rpm, 4 h, 40 °C1:2n-hexane82.014.23.5[284]
Butyl ButyrateCandida rugosa lipase immobilized onto multiwalled carbon nanotubes
11.6% (w/w)
Orbital shaker, 200 rpm, 4 h, 37 °C1:3n-hexane91.513.23.3[285]
Butyl ButyrateCandida rugosa lipase immobilized onto multiwalled carbon nanotubes
4.2% (w/w)
Orbital shaker, 200 rpm, 4 h, 30 °C1:1n-hexane92.713.43.3[175]
Isopentyl ButyrateLipozyme® TL IM
14.7% (w/w)
Orbital shaker, 180 rpm, 6 h, 40 °C,1:2n-hexane96.1259.543.3[286]
Isopentyl ButyrateLipozyme® IM-20
1.9% (w/w)
Orbital shaker, 150 rpm, 24 h, 40 °C1:1n-hexane~92.036.31.51[287]
Isopentyl ButyrateRhizopus oryzae lipase immobilized onto Purolite (PMMA with epoxide and octadecyl groups)
35 U
Orbital shaker, 1200 rpm, 12 h, 30 °C1:2cyclohexane91.059.04.9[108]
Hexyl ButyrateNovozym® 435
13.8 g mol−1
Orbital shaker, 200 rpm, 3 h, 40 °C1:1n-hexane93.080.126.7[273]
Hexyl ButyrateCandida rugosa lipase immobilized onto Diaion HP20
10.8% (w/w)
Orbital shaker, 200 rpm, 3 h, 40 °C1:2n-heptane94.581.827.6[288]
Ethyl ValerateCandida rugosa lipase immobilized onto microemulsion-based organogels
-
Orbital shaker, 150 rpm, 144 h, 40 °C1:1.6cyclohexane99.913.00.1[289]
Ethyl ValerateCandida rugosa lipase immobilized onto app membrane
696.6 U in 20 mL
Screw-capped flasks, 150 rpm, 24 h, 40 °C [a]2:1
1:1
solvent-free
n-hexane
34.8
67.2
164.2
583.2
6.8
24.3
[178]
Ethyl ValerateNovozym® 435
10.06 g L−1
Microwave reactor, 0.67 h, 50 °C1:1solvent-free69.28.012.0[290]
Ethyl ValerateCandida antarctica lipase B
12% (w/v)
Orbital shaker, 1000 rpm, 2 h, 40 °C
Microwave, 100 W, 1 h, 40 °C
1:2
1:2
solvent-free90.0
82.0
518.1
472.1
259.1
472.1
[291]
Ethyl ValerateAspergillus oryzae immobilized onto an electrospun nanofibrous membrane
1.10 g L−1
Magnetical stirring in oil bath, 250 rpm, 24 h, 50 °C [a]1:1solvent-free72.123.31.0[292]
Ethyl ValerateBurkholderia cepacian lipase immobilized onto sodium alginate
20% (w/v)
Orbital shaker, 150 rpm, 48 h, 37 °C1:2n-heptane87.158.61.2[293]
Ethyl ValerateCandida rugosa lipase immobilized onto biogenic sílica/magnetite/
graphene oxide
2.1% (w/w)
Orbital shaker, 200 rpm, 3 h, 40 °C1:2n-heptane90.487.529.1[294]
Pentyl ValerateCandida rugosa lipase immobilized onto nanocrystalline cellulose, silica, and polyethersulfone
3 mg mL−1
Orbital shaker, 200 rpm, 3 h, 50 °C1:2cyclohexane91.3--[295]
Pentyl ValerateCandida rugosa lipase immobilized onto microemulsion-based organogels
70 g L−1
Orbital shaker, 150 rpm, 199 h, 37 °C 2.9:2cyclohexane/AOT99.917.00.1[109]
Ethyl
Hexanoate
Antarctic pseudomonas (AMS8) immobillized onto chitosan
5% (w/w)
Orbital shaker, 200 rpm, 2 h, 20 °C1:1toluene
solvent-free
69.0
47.0
2.5
372.8
1.3
93.2
[296]
[a] Studies that include water addition in the beginning of the reaction.
Table 2. Short-chain ester syntheses from aromatic reagents using immobilized lipases.
Table 2. Short-chain ester syntheses from aromatic reagents using immobilized lipases.
EsterBiocatalyst
Biocat. Loading
Reaction ConditionsAcid:
Alcohol Molar Ratio
SolventConversion (%)Concentration (g L−1)Productivity (g L−1 h−1)Ref.
Phenethyl FormateNovozym® 435
15 g L−1
Orbital shaker, 150 rpm, 4 h, 40 °C1:51,2-dichloroethane95.914.43.6[297]
Phenethyl FormateNovozym® 435
15 g L−1
Orbital shaker, 150 rpm, 30 min, 30 °C1:1n-hexane69.2--[121]
Benzyl PropionateNovozym® 435
10% (w/w)
Thermal bath with mechanical agitation, 24 h, 65 °C1:5solvent-free44.0127.65.3[298]
Benzyl PropionateNovozym® 435
10% (w/w)
Fed-batch system with molecular sieves, 150 rpm, 6 h, 50 °C1:5solvent-free99.0270.745.1[299]
Benzyl ButyratePseudomonas cepacian lipase immobilized onto polyvinyl alcohol and chitosan
8.6% (w/w)
Ultrasound, 100 W, 3 h, 52 °C1:3isooctane40.0--[176]
Benzyl ButyrateNovozym® 435
2% (w/w)
Round bottom flask with fish-clip spinner, 20 h, 52 °C; irreversible esterification reaction (chemical drying)1:1methyl tert-butyl ether82.0160.48.0[300]
Methyl CinnamateCandida sp. lipase immobilized onto ordered mesoporous silicon C8
10 mg
Orbital shaker, 220 rpm, 2 h, 70 °C1:3isooctane90.41.50.8[301]
Table 3. Reusability of different biocatalysts in short-chain ester syntheses.
Table 3. Reusability of different biocatalysts in short-chain ester syntheses.
EsterBiocatalystNumber of Sequential ReactionsRef.
Butyl PropionateNovozym® 43520[275]
Hexyl FormateNovozym® 43520[265]
Ethyl ButyrateCandida rugosa lipase immobilized into polyurethane foams20[278]
Butyl AcetateNovozym® 43514[259]
Octyl FormateNovozym® 43510[264]
Pentyl AcetateLipozyme® 43510[21]
Pentyl ValerateCandida rugosa lipase immobilized onto microemulsion-based organogels10[109]
Ethyl ButyrateCandida antarctica lipase B immobilized onto styrene beads8[183]
Isobutyl PropionateFermase® CALB 100007[277]
Isobutyl PropionateNovozym® 4357[37]
Butyl AcetateCandida antarctica lipase B immobilized onto porous styrene-divinylbenzene beads6[182]
Butyl ButyrateThermomyces lanuginosus lipase immobilized onto glyoxyl-agarose beads5[194]
Butyl AcetateNovozym® 4355[256]
Ethyl ValerateCandida rugosa lipase immobilized onto amine polypropylene membrane5[178]
Ethyl ValerateAspergillus oryzae immobilized onto an electrospun nanofibrous membrane5[292]
Isopentyl ButyrateRhizopus oryzae lipase immobilized onto Purolite (PMMA with epoxide and octadecyl groups)5[108]
Isobutyl ButyrateLipozyme® RM-IM4[259]
Isopentyl AcetateStaphylococcus simulans lipase immobilized onto CaCO34[61]
Isopentyl PropionateNovozym® 4354[254]
Butyl PropionateNovozym® 4353[276]
Butyl AcetateRhyzopus oryzae lipase immobilized onto Celite 5453[210]
Ethyl ValerateNovozym® 4352[290]
Isopentyl AcetateLipozyme® IM-202[266]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sousa, R.R.d.; dos Santos, M.M.; Medeiros, M.W.R.; Manoel, E.A.; Berenguer-Murcia, Á.; Freire, D.M.G.; Fernandez-Lafuente, R.; Ferreira-Leitão, V.S. Immobilized Lipases in the Synthesis of Short-Chain Esters: An Overview of Constraints and Perspectives. Catalysts 2025, 15, 375. https://doi.org/10.3390/catal15040375

AMA Style

Sousa RRd, dos Santos MM, Medeiros MWR, Manoel EA, Berenguer-Murcia Á, Freire DMG, Fernandez-Lafuente R, Ferreira-Leitão VS. Immobilized Lipases in the Synthesis of Short-Chain Esters: An Overview of Constraints and Perspectives. Catalysts. 2025; 15(4):375. https://doi.org/10.3390/catal15040375

Chicago/Turabian Style

Sousa, Ronaldo Rodrigues de, Michelle M. dos Santos, Matheus W. R. Medeiros, Evelin A. Manoel, Ángel Berenguer-Murcia, Denise Maria Guimarães Freire, Roberto Fernandez-Lafuente, and Viridiana Santana Ferreira-Leitão. 2025. "Immobilized Lipases in the Synthesis of Short-Chain Esters: An Overview of Constraints and Perspectives" Catalysts 15, no. 4: 375. https://doi.org/10.3390/catal15040375

APA Style

Sousa, R. R. d., dos Santos, M. M., Medeiros, M. W. R., Manoel, E. A., Berenguer-Murcia, Á., Freire, D. M. G., Fernandez-Lafuente, R., & Ferreira-Leitão, V. S. (2025). Immobilized Lipases in the Synthesis of Short-Chain Esters: An Overview of Constraints and Perspectives. Catalysts, 15(4), 375. https://doi.org/10.3390/catal15040375

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop